174
THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE OCEAN by KEVIN DOUGLAS LEAMAN B.S., University of Michigan (1970) M.S., University of Michigan (1972) SUBMITTED IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY at the MASSACHUSETTS INSTITUTE OF TECHNOLOGY and the WOODS HOLE OCEANOGRAPHIC INSTITUTION June, 1975 Signature of Author........7.- ....... ...... Joint Program in Oceanography, Massachusetts Institute of Technology - Woods Hole Oceanographic Institution, and Department of Earth and Planetary Sciences, and Department of Meteorology, Massachusetts Institute of Technology, June 1975 I I Certified by.... .S .v r;..v; .- r. r ~ . " . . iV w.. *sor SKThesis Supervisor Accepted by.. ..-. . ... . .. ... ., .. . .. ... ChairFmi , Joint Oceanogra y Committee in the Earth Sciences, Massachusetts Institute of Technology - Woods Hole Oceanographic. Institution

THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

THE VERTICAL PROPAGATION OF INERTIAL WAVES

IN THE OCEAN

by

KEVIN DOUGLAS LEAMAN

B.S., University of Michigan(1970)

M.S., University of Michigan(1972)

SUBMITTED IN PARTIAL FULFILLMENT OF THEREQUIREMENTS FOR THE DEGREE OF

DOCTOR OF PHILOSOPHY

at the

MASSACHUSETTS INSTITUTE OF TECHNOLOGY

and the

WOODS HOLE OCEANOGRAPHIC INSTITUTION

June, 1975

Signature of Author........7.- ....... ......Joint Program in Oceanography, MassachusettsInstitute of Technology - Woods HoleOceanographic Institution, and Department ofEarth and Planetary Sciences, and Departmentof Meteorology, Massachusetts Institute ofTechnology, June 1975

I I

Certified by.... .S .v r;..v; .- r. r ~ . " . . iV w.. *sor

SKThesis Supervisor

Accepted by.. ..-.. ... . .. ... ., .. . .. ...ChairFmi , Joint Oceanogra y Committee in theEarth Sciences, Massachusetts Institute ofTechnology - Woods Hole Oceanographic.Institution

Page 2: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

2

THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE OCEAN

by

Kevin Douglas Leaman

Submitted to the Massachusetts Institute of Technology-Woods HoleOceanographic Institution Joint Program in Oceanography in June,1975, in partial fulfillment of the requirements for the degreeof Doctor of Philosophy.

ABSTRACT

A set of vertical profiles of horizontal ocean currents, ob-tained by electro-magnetic profilers in the Atlantic Ocean southwestof Bermuda in the spring of 1973, has been analyzed in order to studythe vertical structure and temporal behavior of internal waves, par-ticularly those with periods near the local inertial period. An im-portant feature of the observed structure is the polarization of hor-izontal velocity components in the vertical. This polarization, alongwith temporal changes of the vertical wave structure seen in a timeseries of profiles made at one location, has been related to the di-rection of vertical energy flux due to the observed waves. Whereasthe observed vertical phase propagation can be affected by horizontaladvection of waves past the point of observation, the use of wave po-larization to infer the direction of vertical energy propagation hasthe advantage that it is not influenced by horizontal advection. Theresult shows that at a location where profiles were obtained oversmooth topography, the net energy flux was downward, indicating thatthe energy sources for these waves were located at or near the seasurface. An estimate of the net, downward energy flux (- .2 - .3erg/cm2/sec) has been obtained. Calculations have been made whichshow that a frictional bottom boundary layer can be an important en-ergy sink for near-inertial waves. A rough estimate suggests thatthe observed, net, downward energy flux could be accounted for byenergy losses in this frictional boundary layer. A reflectioncoefficient for the observed waves as they reflect off the bottomhas been estimated.

In contrast, some profiles made over a region of rough topog-raphy indicate that the rough bottom may also be acting to generatenear-inertial waves which propagate energy upward.

Calculations of vertical flux of horizontal kinetic energy,using an empirical form for the energy spectrum of internal waves,show that this vertical flux reaches a maximum for frequencies 10% -20% greater than the local inertial frequency. Comparison with pro-filer velocity data and frequency spectra supports the conclusionthat the dominant waves had frequencies 10% - 20% greater than theinertial frequency. The fact that the waves were propagating energyin the vertical is proposed as the reason for the observed frequencyshift.

Page 3: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

Finally, energy spectra in vertical wave number have beencalculated from the profiles in order to compare the data with anempirical model of the energy density spectrum for internal wavesproposed by C. Garrett and W. Munk (1975). The result shows thatalthough the general shape and magnitude of the observed spectrumcompares well with the empirical model, the two-sided spectrum isnot symmetric in vertical wave number. This asymmetry has beenused to infer that more energy was propagating downward than upward.These calculations have also been used to obtain the coherence be-tween profiles made at the same location, but separated in time(the so-called dropped, lagged, rotary coherence). This coherenceis compared with the aforementioned empirical model. The coherenceresults show that the contribution of the semidiurnal tide to theenergy of the profiles is restricted to long vertical wave lengths.

Thesis Supervisor: Dr. Thomas B. SanfordTitle: Associate Scientist

Page 4: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

ACKNOWLEDGEMENTS

I wish to thank above all Dr. Thomas B. Sanford of the

Woods Hole Oceanographic Institution for providing the data on

which this report is based, and for many helpful comments and

suggestions. Initial acquisition and processing of data obtained

by the electro-magnetic profiler was successful primarily due to

the efforts of R. G. Drever, E. Denton, A. Bartlett, J. Dunlap

and J. Stratton, all of the above institution. Support for the

experiment which is described in this report was provided by the

Office of Naval Research under contracts N00014-66-C-0241, NR 083-

004 and N000 14-74-C-0262, NR 083-004.

For my personal support as a graduate student I thank the

National Science Foundation for its award to me of a graduate

fellowship for three years. I also thank the Office of Naval

Research (under the above contracts) for providing part of my

support.

Page 5: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

TABLE OF CONTENTS

Page

Abstract

Acknowledgements

List of Figures

Chapter I Introduction and Historical Review

Chapter II Description and Analysis of Data from theFive-day Time Series of Profiles

a.) Dominance of inertial-period motions inEMVP profiles--the "mirror-imaging" of.profiles

b.) Low-frequency "geostrophic" profiles

c.) High-frequency velocity components inthe upper 2.5 km as a function of time

d.) The average perturbation kinetic energyprofile, and a smoothed profile of Brunt-V*is*l* frequency

e.) The stretching of the vertical coordinateand the normalization of velocity by themean Brunt-Visal profile

f.) Spectral decomposition of stretched pro-files

g.) Paired profiles

Chapter III Vertical Polarization and the Vertical Propa-gation of Internal-inertial Waves

a.) Summary of data

b.) The vertical spatial behavior of hori-zontal internal wave velocity components

10

21

21

23

27

32

36

42

54

57

57

58

Page 6: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

Chapter III

Chapter IV

Chapter V

Chapter VI

Chapter VII

Appendix A

Appendix B

Appendix C

Appendix D

Bibliography

c.) Polarization in the deep water

d.) Comparison with meteorological observa-tions

The Use of the Ratio of Clockwise to Counter-clockwise Energy as a Reflection Coefficient.Reflection of Internal Waves from a SmoothBottom

a.) Calculation of a reflection coefficientfor the observed waves

.b.) A model for the reflection of internalwaves from a smooth bottom

Comparison of Vertical Wave Number Spectrawith Theoretical Models. Computation ofDropped, Time-lagged, Rotary Coherence (DLRC).The Vertical Energy Flux of Internal-inertialWaves

a.) Vertical wave number spectra

b.) Dropped, time-lagged, rotary coherenceof profiles

c.) The vertical energy flux of near-inertialwaves

Some Observations over Rough Topography

Conclusions, a Further Discussion of Some ofthe Data and Recommendations for FutureExperiments

Profile Locations and Times

Use of Electro-magnetic Velocity Profilersto Measure Internal Waves

Accuracy and Precision of Observations

A More General Treatment of the Partition ofEnergy Between Clockwise and CounterclockwiseSpectra

Biographical Note

Page71

89

98

107

119

126

142

146

159

163

170

174

Page 7: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

LIST OF FIGURES

FigureNumber Page

1. Electro-magnetic velocity profiler 16

2. Bathymetry of the MODE region. Location of central 18mooring time series is shown by an X. The ridge overwhich rough topography profiles were made is circled.

3. Paired drops, 219D and 221D, made at the same loca- 22tion and with a time separation of one half of aninertial period.

4. Average profile of the east velocity component during 25the five-day time series.

5. Average profile of the north velocity component during 26the five-day time series.

6. Contours of the east velocity component during the 28five-day time series. Negative east components areindicated by hatched regions.

7. Contours of the north velocity component during the 29five-day time series. Negative north components areindicated by hatched regions.

8. Average profiles of high-frequency horizontal kinetic 33energy and Brunt-VaislIN frequency.

9. Stretched and unstretched vertical pressure coordinates. 39

10. East component of profile 219D as originally obtained 40(a), after removal of the mean profile (b), and afterstretching and normalizing(c).

11. North component of profile 219D as originally obtained 41(a), after removal of the mean profile (b), and afterstretching and normalizing (c).

12. Total spectrum (east and north energy) for 9 down pro- 43files.

Page 8: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

FigureNumber Page

13. Clockwise (CW) and counterclockwise (ACW) spectra for 459 down profiles. In the text, CW is C(m') and ACW isA(m').

14. Spectra of east and north velocity components for 9 48down profiles.

15. Total spectrum (east and north energy) for 9 up pro- 49files.

16. Clockwise (CW) and counterclockwise (ACW) spectra 50for 9 up profiles.

17. Spectra of east and north velocity components for 9 51up profiles.

18. Simultaneous paired drops, 235U and 236U, separated 55horizontally by 4.8 km.

19. Horizontal coordinate rotation (see equations 111-2). 59

20. Direction of rotation of the horizontal velocity vec- 59tor of an internal wave for positive and negativevertical wave numbers. The curved arrow shows thedirection in which V rotates with increasing z.

21. Propagation diagram for internal waves. Cg is the 62group velocity vector, k is the wave number vector,and V is a vector representing water velocity.

22. Reflection coefficient A(m')/C(m'). 79

23. Comparison of theoretical and observed total energy 96spectra in stretched vertical wave number.

24. Dropped, lagged, rotary cospectrum (normalized) be- 101tween velocity components U1 and Us (see text for afurther description of Figures 24 to 27).

25. Dropped, lagged, rotary cospectrum (normalized) be- 102tween velocity components VI and V3 .

26. Dropped, lagged, rotary quadrature spectrum (normal- 103ized) between velocity components U, and U,.

Page 9: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

FigureNumber Page

27. Dropped, lagged, rotary quadrature spectrum (normal- 104ized) between velocity components V1 and V3.

28. Frequency dependence of the vertical flux of hori- 114zontal kinetic energy, using a theoretical form forthe energy spectrum of the internal wave field.

29. Frequency spectrum, obtained by averaging 12 spectra 117computed from current meter records at 1500 m inthe MODE region. Energy density units are(cm/sec)2/cph.

30. Smoothed hodograph of profile 190D between 3000 dbar 122and 4500 dbar. Depths are in hundreds of decibars.

31. Smoothed hodograph of profile 192D between 3000 dbar 123and 4500 dbar. Depths are in hundreds of decibars.

Page 10: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

Chapter I Introduction and Historical Review

For many years the subject of internal waves has been of in-

tense interest to meteorologists, aeronomists, and oceanographers.

These internal waves are oscillations in the atmosphere or ocean

that are supported by buoyancy and pressure forces, and also by the

fictitious Coriolis force introduced by the rotation of the earth.

Oscillations having time and length scales appropriate to internal

waves have been observed at many locations in the ocean. In the

atmosphere, motions having time and space scales appropriate to

internal waves have also been observed. However, until recently,

meteorologists and oceanographers have been studying internal wave

phenomena from decidedly different points of view. This has been

brought about primarily by the different types of instrumentation

used in the two disciplines. Observations of atmospheric internal

waves are discussed first.

In the years since World War II, several methods have been

developed by meteorologists for observing variables such as wind

and temperature in the atmosphere. Wind measurements, which are

of primary interest in the present study, have been performed using

the following methods: tracking the distortion of smoke (actually

sodium) trails left by high altitude rockets; tracking of radar re-

flecting balloons with ground-based radar; visual tracking of meteor

trails by telescopes; tracking chaff (strips of aluminized reflect-

ing material) with ground-based radar; and radio-echo tracking of

Page 11: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

meteor trails. All of these methods have their disadvantages.

Rockets, for example, are expensive to use, and they have not

usually been applied to rapidly repeated measurements in the at-

mosphere. Visual and radio-echo tracking of meteors depend, of

course, on a sufficient number of meteors entering the upper at-

mosphere. Except in rare cases when meteor showers occur, the

number of meteors in the atmosphere has usually been insufficient

to study high-frequency oscillations in the wind field, such as

internal waves. More recently, tracking of ascending, radar re-

flecting balloons has provided better measurements of atmospheric

winds.

All of these methods, although diverse in concept, have in

common the fact that they provide profiles of horizontal atmos-

pheric winds as a function of altitude. Many examples of these

profiles have been presented in the literature (see, for example,

the paper by L. Broglio in "Proceedings of the First International

Symposium on Rocket and Satellite Meteorology," in which results

from rocket profiles are presented, and the radar balloon observa-

tions of Endlich, Singleton and Kaufman (1969)). Thus, the atten-

tion of meteorologists and aeronomists has historically been di-

rected toward observing the vertical structure of the atmospheric

wind field. Attempts at using repeated profiles to determine the

amount of energy contributed by high-frequency (internal wave and

turbulent) oscillations at different time scales or frequencies

have been notably less successful than the measurements of the

Page 12: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

profiles themselves. These attempts have usually consisted of try-

ing to fit diurnal and semidiurnal period sinusoids to the winds

observed at a given altitude (Elford and Robertson, 1953; Smith,

1960).

The situation in the oceans has been substantially different.

Until recently, the primary tools which oceanographers have used to

observe oscillations with frequencies in the internal wave range

have been fixed-location current meters, temperature sensors, and

neutrally buoyant floats. The current meter allows one to measure,

at a fixed point in the ocean, the horizontal current vector as a

function of time. If current meters are placed at many locations

in a part of the ocean (by being put on moorings), then time meas-

urements in a spatial array are obtained. These arrays allow one

to obtain good temporal resolution of velocity, but only at a

limited number of points in space.

Current meters have provided a large volume of data on the

temporal behavior of the ocean. Fofonoff (1969) showed that the

behavior of horizontal velocity components and temperature in the

internal wave frequency band (those frequencies corresponding to

periods between one half of the local pendulum day, 1/(2sin(Lat.))

days, and the local stable oscillation period of a particle of

water, or the Brunt-VIisali period) could be reasonably well re-

lated to the expected behavior of internal waves in a linear model.

Webster (1968) summarizes many of the current meter observations

that have been used to study internal waves and other high-frequency

Page 13: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

oceanic phenomena. Most of these current meter records have shown

that the dominant contributions to horizontal kinetic energy spec-

tra come from frequencies near the local inertial frequency, cor-

responding to the aforementioned period of one half of a pendulum

day, and from the tides.

Temperature observations have been made using several dif-

ferent methods. Hauritz, Stommel and Munk (1959) observed temp-

erature oscillations as a function of time at two fixed depths,

50 and 550 meters, near Bermuda. These data yielded frequency

spectra for the temperature oscillations.

Another method, employed by Charnock (1965) and LaFond and

LaFond (1971), is to tow a weighted thermistor chain behind a

vessel. If the speed of the vessel is large compared to the hori-

zontal phase propagation speed of the waves being observed, spec-

tral analysis of the temperature trace obtained at a gi.ven depth

gives a horizontal wave number spectrum of the observed isotherm

displacements. Similar observations can be obtained by towing an

isotherm-following "fish" behind a research vessel (Katz, 1973,

1975).

The use of neutrally buoyant floats has also provided data

on internal waves. Pochapsky (1972) used such floats to observe

oscillations of temperature and horizontal velocity with periods

in the internal wave range. Voorhis (1968) obtained measurements

of vertical motion with a neutrally buoyant float which rotates in

the presence of vertical flow past the float. The rate of rota-

tion of the float gives an estimate of this vertical flow.

Page 14: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

Many of the observations described above (as well as others

which have not been mentioned) have been gathered together by

Garrett and Munk (1972, 1975) in an attempt to model the energy

spectrum of internal waves in frequency and horizontal wave number.

The model depends not only on individual spectra of temperature or

velocity, but also on the correlation, or coherence, between obser-

vations made at different points of a spatial array. Coherence

observations have also been gathered together and described by

those authors (Garrett and Munk, 1972, 1975).

A major difficulty with this type of model construction has

been lack of knowledge concerning the vertical structure of the

wave field. This vertical structure is difficult to determine

using velocity (or temperature) measurements from moorings, since

only a limited number of instruments can be placed on any one

mooring. Thus, vertical profiles of horizontal velocity, which

have been the most common method of observation in the atmosphere,

have been the least common type of observation in oceanographic

work.

Recently, however, several methods have been developed that

allow the oceanographer to obtain vertical profiles of horizontal

current in the ocean that are analogous to profiles obtained in the

atmosphere. One such method involves the use of acoustic trans-

ponders located on the ocean bottom to track a float as it descends

through the water column (Rossby, 1969; Pochapsky and Malone, 1972).

Another method uses a moored vertical cable with inclinom-

eters located at various points along the cable (McNary, 1968).

Page 15: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

Horizontal currents cause the cable to distort. This distortion,

as measured by the inclinometers, can then be converted to values

of the horizontal velocity component along the cable. A third

method uses horizontal electrical currents induced by the flow of

ocean water through the terrestrial magnetic field (Sanford, 1971).

The device which measures these electrical currents is called an

electro-magnetic velocity profiler, or EMVP. A picture of this

device, developed by T. Sanford and R. Drever of the Woods Hole

Oceanographic Institution, is shown in Figure 1. The EMVP is re-

leased from a research vessel, and as it falls to the ocean bottom

it continuously records the horizontal component of electrical

current, as well as temperature, conductivity, and pressure. When

the EMVP reaches the bottom, it reverses and ascends through the

water column, again recording the above variables. There are no

hard connections (cables, etc.) between the research vessel and

the EMVP: the only communication from EMVP to ship is through

acoustic telemetry. All data are recorded on 7-track magnetic

tape within the EMVP. Upon return to the ship, the data on the

tape is reformatted onto 9-track tape, and several computer programs

are used to convert the internally recorded variables to their

physical counterparts, such as temperature. Another routine, using

a conversion equation, converts horizontal electrical current den-

sity to horizontal water flow relative to an unknown, but depth

independent, horizontal velocity. A further description of the

EMVP and its operation can be found in Sanford, et. al. (1974).

Page 16: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

'I

/

7'r

J 4.;

'

4 -

F'

ii

~;

47 ~

?

V45

4

1

//

I1~

II

I ~44 X

1'

4WA

J

I V

V.'

V.

~,

-I

'4.4

'(544')

p~

j'i

I~

;d

Shflfl.S

bih

&.L

A

iJa.a

.i.~

>

~

4,1wo

xtI ft

~ .J

na

4-0r-(U

4-,

VA

CL

i -.

I-4F

IR

'I 7-7

1771T

-77R

",,

Page 17: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

The horizontal velocity component is not the only quantity

that has been measured by vertical profiling. Vertical profiles

of temperature, conductivity and salinity have been obtained by

raising and lowering a CTD (conductivity-temperature-depth sensor).

Such data have been presented by Hayes (1975) and Hayes, et. al.

(1975).

We turn now to a description of the experiment which is

discussed in this report. In the spring of 1973 a large-scale

oceanographic experiment called MODE (Mid-Ocean Dynamics Experi-

ment) was carried out in the Atlantic Ocean, southwest of Bermuda.

A map showing the bathymetry in the region of this experiment is

shown in Figure 2. The EMVP was used to take a large number of

vertical profiles during the experiment. The locations, dates and

times of all profiles are given in Appendix A. We are going to

concentrate particularly on two subsets of all the profiles. The

first subset consists of a time series of repeated profiles made

at the center of the experiment (69* 40' W, 27* 59' N). This will

be called the "central mooring time series," since the central

mooring of the MODE experiment was located near these coordinates.

This time series, made up of 20 profiles, lasted from June 11-15,

1973. (These profiles are designated by an asterisk in Appendix A).

As can be seen from Figure 2, the location of the series is over a

region of smooth bottom topography. The second subset is made up

of some profiles taken around a ridge east of the time series, in a

region of rough topography. These locations are shown in Figure 2.

Page 18: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

Al i I ~I .~I I . I a I30

AMES indicate seamounts MODE Region Bathymetry byand minimounts

) .f ,, P, Bush)epths in corrected meters o: (

AOML/NOAA/ _ -29

4I /' 'VANGUARDS;VOLKSTWAMANN .

O MODE 44602

-28

G UL .SWALLOW "-...L",

26

Figure 2. Bathymetry of the MODE region. Location of centralmooring time series is shown by an X. The ridge overwhich rough topography profiles were made is circled.

Page 19: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

Any profile that has ever been made by the EMVP has shown

two dominant contributions to the velocity structure in the verti-

cal. The first contribution is a low-frequency part, which does

not change appreciably from one profile to the next at the same

location, and which corresponds well with current shear derived by

geostrophic shear calculations in that region. The second part is

a high-frequency contribution which is evidently due to internal

waves, particularly those with frequencies near the local inertial

frequency. It is this high-frequency part that will be of main

interest in what follows.

Chapter II gives a description and analysis of the velocity

data obtained fr',m the five-day time series of profiles. Plots of

mean flow and average perturbation horizontal kinetic energy versus

depth are given, as are contour plots showing the evolution of

features in the velocity field in time, and spectra of 'the high-

frequency profile structure in vertical wave number.

One of the most interesting results obtained from the spec-

tral analysis in Chapter II is that there is a definite elliptical

polarization of the high-frequency waves in the profiles. That is,

the horizontal current vector in a profile (after the mean has been

removed) clearly tends to rotate with depth. This polarization and

its relation to the vertical propagation of internal wave energy is

described in Chapter III. In Chapter III, comparisons will be made

between our observations and similar observations made in the atmos-

phere.

Page 20: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

Chapter IV describes the reflection of internal waves from

a smooth bottom. The spectra presented in Chapter II are used to

attempt to calculate reflection coefficients for these waves.

Chapter V presents a comparison of the vertical wave number

spectra in Chapter II to the theoretical models of Garrett and

Munk (1972, 1975). The coherence of velocity components between

pairs of profiles separated by a time lag is also presented. Ver-

tical energy flux of the observed waves is discussed in more detail.

Chapter VI discusses the profiles obtained over rough topog-

raphy. Chapter VII provides some conclusions, a further discussion

of some of the data, and recommendations for future experiments.

~Y_~XIL________IY_____ _~1_1 I___ ~I .~

Page 21: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

Chapter II Description and Analysis of .Data from the Five-day

Time Series of Profiles

It was mentioned in the introduction that EMVP data is his-

torically more akin to meteorological measurements in the upper at-

mosphere than to oceanographic data. In particular, profiler data

emphasizes the vertical structure of the ocean or atmosphere, while

until recently most oceanographic work has emphasized the temporal

behavior of oceanic motions. This chapter will present, in various

forms, horizontal velocity data obtained during the five-day series

of profiles, and will describe some of the methods used to reduce

and analyze this data. Succeeding chapters are devoted to the

interpretation of results obtained from this analysis. The reader

may refer to Appendix B for a brief discussion of internal wave

measurements made by the EMVP and to Appendix C for a discussion of

the accuracy and precision of data presented in the following sec-

tions.

a.) Dominance of inertial-period motions in EMVP profiles--the"mirror-imaging" of profiles

Figure 3 shows the east and north components of two profiles,

219D and 221D. (The letter D signifies that the down portion of a

drop was used. The letter U will be used to denote the up portion

of a drop.) The time separation between the two profiles is approx-

imately one half of an inertial period. The vertical sampling rate

is one point every 10 dbar. It is evident from the figure that,

~ 11~1111~- 1_1_L~L1_-i (--III--L~-._L~1-~i-~__l;_l_ ___-QI~ ilY

Page 22: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

219D & 221D

EAST5 (CM/SI) NORTH, (CM/S)

Paired drops, 219D and 221D, made at the same locationand with a time separation of one half of an inertialperiod.

Cr)U4l

Figure 3.

VELOCITY PROFILES

Page 23: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

superimposed on a low-frequency shear in the profiles, there is a

considerable amount of high-frequency activity. This high-frequen-

cy activity is evident in the almost total reversal of the velocity

structure over the time interval. If we remove the low-frequency

shear of the profiles (by averaging all drops in the time series

and subtracting the low-frequency shear profile from each drop sep-

arately), we find that the high-frequency part that is left over

is negatively correlated in profiles made one half of an inertial

period apart. This negative correlation, or "mirror-imaging," is

an indication that the EMVP profiles are strongly dominated in the

high frequencies by motions with periods close to the local iner-

tial period. It will be shown later that some of the observed

variability is also contributed by the semidiurnal tide.

b.) Low-frequency "geostrophic" profiles

For ease of discussion we will call oscillations with fre-

quencies less than the local inertial frequency "low-frequency"

oscillations, or the "geostrophic" part of the flow. Similarly,

oscillations with frequencies greater than or equal to the inertial

frequency will be called "high-frequency," or "internal wave"

oscillations. It should be clear that these terms are used solely

to simplify the description of the data. We do not intend to

imply that all motions with frequencies less than inertial are in

geostrophic balance. Also, high-frequency motions are not necessar-

ily due to internal waves. A certain amount of high-frequency en-

ergy may be contributed by turbulence, for example. It is known,

H~^-~" ~L I^~" ~'-Y~--eXlf~L~ ~-LII IIC-.-~C-~CY---

Page 24: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

however, that there is a "well,." or depression in frequency spectra

obtained in the MODE region (see Figure 29). This well exists in

the range of periods from the inertial period to periods of several

days. Therefore, averaging velocity data over several days will

partition the profile data into high-frequency and low-frequency

parts. In order to characterize more fully this partition of en-

ergy between the low-frequency "geostrophic" and high-frequency

"internal wave" parts of the profiles, we have averaged all 20 pro-

files in the time series. (The 20 profile time series occupies a

time interval of about 103 hours.) This averaging was done simply

by removing the "barotropic," or depth-averaged, part of each pro-

file, and then taking an average over the 20 drops at each level.

The depth-averaged part of each profile was removed because it is

unknown and is not the true barotropic part of the velocity profile

(see Sanford (1971)). The resulting low-frequency profiles are

shown in Figures 4 and 5 for the east and north components.

There is some indication in these mean profiles of a decrease

in the mean shear in the 18* water (around 400 dbar). There is

also an indication that the mean shear increases again near the

bottom, in the Antarctic Bottom Water. In terms of the geostrophic

shear relation this might be expected. Figure 8 shows the mean

Brunt-Vaisal profile at the location of the time series. Since

the Brunt-Viisala frequency is proportional to the square root of

the vertical potential density gradient, this profile shows that

the vertical density gradient has a relative minimum at about

~4~ III_~YI__I^_______1II__)__ -~ll ~ l~ ~ -LI~ si-L

Page 25: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

5- DA Y SERIES

AVERAGE EAST PROFILE

-30 -20 -10 01000 -

10001

2000

3000

4000

5000

6000

(CM/SEC)

20

Figure 4. Average profile of the east velocity componentduring the five-day time series.

30

__ L__~ rl__~_Y)___I-Y ~ I--1PII~YL~U

Page 26: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

26.

5- DAY SERIES

AVERAGE NORTH PROFILE (CM/SEC)

-30 -20 -10 0 10 20 300 1 I 1 I

I I I I I I I I I

Average profile of the north velocity componentduring the five-day time series.

1000 -

2000 -

3000 -

4000 1

5000-

6000

Figure 5.

~--_-~-e-U-iiPYl~i ~- _III--.L.^PILYL1-II~ --LIIIL.. ~~~IL~-

Page 27: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

400 dbar, and increases slowly from about 4500 dbar to the bottom.

The vertical shear of a geostrophic flow is proportional to the

horizontal density gradient. It is reasonable to expect that the

shear flow has caused the density field to distort in such a manner

that the horizontal density gradient at a point is proportional to

the vertical density gradient at the same point. Keeping in mind

the fact that we have assumed the low-frequency flow to be in geo-

strophic balance, the above would seem to be a reasonable explana-

tion of the observed decrease in shear of the mean flow at 400 dbar

and of the increase in this shear as the bottom is approached.

c.) High-frequency velocity components in the upper 2.5 km as afunction of time

From the 20 high-frequency profiles in the five-day time

series, (that is, the profiles obtained after the mean velocities

of Figures 4 and 5 have been removed), contour plots of east and

north velocity components have been made (Figures 6 and 7). These

plots show how various structures in the horizontal velocity field

evolve as time progresses. This evolution appears as changes in

the depth and size of individual features. Before discussing the

significance of these plots, a few comments on their construction

should be made.

First, the distribution of profiles in time is not uniform.

In particular, there are relatively large gaps in time between 224D

and 226D, and between 228D and 230D. Between these two pairs of

drops the profiler had to be used for inter-comparison drops with

Page 28: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

CONTOURS OF ZERO EAST VELOCITY COMPONENT

TIME HOUR oTIME

DATE

DROP o

ONUMBER

500

lu

(I)

1000

1500

2000

2500

8.0

000012-VI-73

000013-VI-73

000014-VI-73

1000 0 0 00000

15-VI-73

( 0 N 3 a N 3 O O N 0 NO - NM to W I-M 0 t ) tonMM M n ) fn n ) qIICM (M w CQ< ~ ~ CY CY Cq( C CY y C

Figure 6. Contours of the east velocity component during thefive-day time series. Negative east components are

indicated by hatched regions.

~~Ll~i i ___ I _ -~--illl~WILLIX-- ICI. .-. -_ .~~-~-1L--IP~ LILC~PLY II.

Page 29: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

CONTOURS OF ZERO NORTH

60oo00oo

13-VI-73

800000

14-VI-73

1000000

15-VI-73

Contours of the north velocity component during thefive-day time series. Negative north componentsare indicated by hatched regions.

TIME HOUR oDATE 0000

12-VI-73

DROPNUMBER

500

1000

1500

2000

2500Figure 7.

__

VELOC/TY COMPONENT

Page 30: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

other vertical profilers, and these inter-comparison drops were not

done at the central mooring. Hence, gaps in the time series re-

sulted. Also, the time interval between profiles in the second

half of the series (after 230D) is generally less than it is for

profiles in the first half. This uneven time distribution allows

some regions of the plots to be contoured more easily than others.

Second, only the 0 cm/sec contour line is shown. Since the

main intent in drawing the contour plots was to attempt to see

vertical phase propagation of the near-inertial waves, a more de-

tailed contouring was not required. Regions with negative velocity

components are indicated by oblique hatching in Figures 6 and 7.

Finally, features for which contouring was ambiguous are

denoted by dashed-line contours. There are two main reasons for

this ambiguity. One is that as the depth increases, the overall

energy level of the profiles decreases (see Figure 8). Thus,

at depths of 2000 dbar to 2500 dbar, the true behavior of the

zero-crossing lines becomes clear. Below 2500 dbar it is very

difficult to draw contours with any confidence. The other reason

is that, even at shallow depths, certain profiles have regions

where a velocity component is close to zero without being clearly

positive or negative. In cases where it appeared that structures

on either side of this ambiguous region were clearly negative,

dashed-line contours were used to join these negative-velocity-

component features together. In other cases, there are regions on

one drop which are clearly positive, while adjacent drops have

II1)L-~ --IPI~PU~ 1\_CILIL--^IIXp_ _ I*CII

Page 31: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

negative structures at the same depth. Such a case occurs at 500

dbar, around drop 242D in the contour plot for the north component.

In these cases, no attempt was made to join adjacent negative-vel-

ocity-component features. Taking note of these difficulties, we

can now point out some of the significant features in these plots.

Perhaps the most striking characteristic is the tendency for

regions of negative velocity to move upward in time. If we assume

that we can connect contours through regions where a velocity com-

ponent is not clearly positive or negative, then most of the struc-

tures with negative velocity can be followed across much of the 103

hours over which the time series lasted. If, on the other hand, we

assume that this connecting of structures is not valid, these nega-

tive-velocity regions will be divided into a series of "hot-dogs,"

with gaps of weakly positive velocity in between. In either case,

the upward motion is evident.

In a few cases, features do appear to move downward in time.

An example occurs at about 1000 dbar in the north plot, around drop

2280. It should be noted that this occurs in a region where the

profiles are widely spaced in time, and the contouring there is not

clear.

Several other points can be made concerning these two fig-

ures. First, if we follow the negative-velocity-component struc-

tures as they move upward in time, we see that, at a given depth,

it usually takes about a day for two successive negative-velocity

zones to pass through a given depth. This indicates that these

_I1__IDIYn____I__I__ ._ .-_lr---- -----..

Page 32: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

structures have a period of about a day, which is near the local

inertial period at this latitude.

Second, if Figures 6 and 7 are overlaid, it is found that,

in general, negative-velocity-component zones in the east plots

occur at depths which are somewhat greater than the depths for the

corresponding zones in the north plots. In other words, the east

and north velocity components are not in-phase with respect to

depth.

Third, the speed with which the negative-velocity-component

zones move upward is about 2 mm/sec at 500 dbar, and this speed in-

creases to about 6 mm/sec at 2000 dbar. We will consider these ob-

servations in more detail in later chapters.

d.) The average perturbation kinetic energy profile, and a smoothedprofile of Brunt-V'isila frequency

The 20 high-frequency profiles (with vertical averages and

the low-frequency or "geostrophic" shear flow removed) were then

used to calculate profiles of high-frequency, or "perturbation,"

kinetic energy. This was done for each profile simply by squaring

the horizontal velocity values at each depth. These horizontal

kinetic energy profiles were then averaged, at each depth, over the

20 drops and a mean kinetic energy profile resulted. This profile

is shown in Figure 8. To simplify plotting, this mean, high-fre-

quency, horizontal kinetic energy profile was averaged in the ver-

tical by taking centered averages 50 dbar long around the points

50 dbar, 100 dbar, 150 dbar, and so on. Plotted along with the

-1--~ - -----I1'~1-~I^(-1~I*L.I1_II.-IIPL *~~I ii-_~_

Page 33: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

MEAN PERTURBATION

KINETIC ENERGY (ERGS/CM 3)

0 I 2 3 4 5 6 7 8

N (CPH)Average profiles of high-frequency horizontalkinetic energy and Brunt-Vi1sa*l frequency.

1000

50

(()Z

u(.)

144(K'44

2000

3000

4000

5000

6000

Figure 8.

1PY-1L-~- -ltm... ~pn^-- II~~ .IXiL~LI~-- IC- .~PII*P-

ENERGYKINETIC

Page 34: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

34

mean horizontal perturbation kinetic energy profile of Figure 8 is

a smoothed Brunt-Vaisala frequency profile, obtained from CTD (con-

ductivity-temperature-depth) stations made at roughly the same time

and location as the time series. Because of difficulties with

measuring velocities by EMVP near the surface (to depths of about

70 dbar), the points at 0 dbar and 50 dbar, showing rather high

kinetic energies, are probably not accurate.

It is possible to calculate an error bar for Figure 8. We

assume: the variance of the noise signal is 0.25 cm2/sec 2 for

either velocity component (Appendix C); noise at a given depth is

uncorrelated with the signal and is uncorrelated between profiles,

but the noise of points in the 50 dbar averages may be correlated

as may the noise in the east and north velocity components. The

contribution of the noise (if it is taussian) will then, for each

point in Figure 8, be distributed as a Chi-square random variable

with (at least) 19 degrees of freedom (Jenkins and Watts, 1969).

The result of applying the Chi-square distribution law is that the

difference between the measured and actual kinetic energy may be

expected to be between .18 and .43 erg/cm 3 95% of the time (the

measured kinetic energy being greater than the actual kinetic en-

ergy). This error estimate has not been plotted in Figure 8, since

it is rather difficult to see (it is about three times the width of

the line depicting the kinetic energy profile). We emphasize that

this error estimate only takes into account instrumental noise. It

does not take into account other sources of error, such as

Page 35: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

35

contamination of the high-frequency profiles due to leakage of

high-frequency energy into the low-frequency profiles (Figures 4

and 5).

We see from Figure 8 that the profiles of mean horizontal

perturbation kinetic energy and mean Brunt-ViisilS frequency are

very similar. The horizontal kinetic energy decreases from the

surface down to about 400 dbar, increases in the main thermocline

to about 1000 dbar, and then steadily decreases down to about 4500

dbar. This parallels the behavior of the Brunt-Vais'li frequency.

It is interesting to note that while the Brunt-Vaisl frequency

slowly increases from about 4500 dbar to the bottom, the mean ki-

netic energy profile increases similarly to about 100-150 dbar off

the bottom, but then decreases again. This near-bottom decrease of

energy may be indicative of the presence of a bottom boundary layer.

The possible existence of a bottom boundary layer will be considered

in more detail in Chapter IV.

We recall that the time series lasted for only 103 hours.

This means that the mean kinetic energy profile was obtained by

averaging over approximately four wave cycles for the inertial

waves. It is not surprising, then, that there is considerable

structure in the mean kinetic energy profile. Presumably, if the

time series of profiles were of longer duration, the average per-

turbation horizontal kinetic energy profile would follow the mean

Brunt-Viisala profile more closely, particularly in the main

thermocline.

Page 36: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

e.) The stretching of the vertical coordinate and the normaliza-tion of velocity by the mean Brunt-Vais3l8 profile

Figures 6 and 7 suggest that the high-frequency motions ob-

served are dominated by internal waves of near-inertial period

which are moving vertically through the water column. An important

quantity which could be calculated from the high-frequency profiles

is the horizontal kinetic energy spectrum as a function of vertical

wave number. However, the data on which we would like to perform

these calculations (that data being the ensemble of profiles in the

time series) is clearly a nonhomogeneous process. Figure 8 shows

that the overall variance (or horizontal perturbation kinetic en-

ergy) of the high-frequency profiles is a function of depth. Also,

an examination of Figure 3 shows that the most energetic waves have

vertical length scales which appear to be shorter in the thermo-

cline than in the deep water. This means that the statistical

properties of the wave field are functions of the Brunt-V~isal

frequency, which in turn is a function of pressure. Some method

must be found which will at least partially convert the process

represented by the ensemble of profiles into a homogeneous one.

Figure 8 gives an idea of the method to be used. Since the hori-

zontal kinetic energy tends to follow the mean Brunt-Vaisila pro-

file, we could make the variance more uniform with depth by normal-

izing the horizontal velocity at a given depth with the square root

of the Brunt-Vaisala* frequency at that depth. But this type of

normalization is what would be expected if the waves were obeying a

WKB approximation (Phillips, 1966), since in this approximation the

Page 37: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

horizontal velocity components of an internal wave are proportional

to (N)1/2, where N is the local Brunt-Vaisla frequency. There-

fore, we chose to normalize the high-frequency velocity profiles

using:. - 1/2

u (z) = u(z)/(N(z)/N ) II-1

where u is an original velocity component, u is a normalized ve-

locity component (both functions of pressure, I), N(z) is the

Brunt-Vaisgli frequency at pressure z and N0 is a reference Brunt-

VMisMl frequency, o = 3 cph. No has been given this value to

facilitate comparison with theoretical vertical wave number spec-

tra presented later (see Chapter V).

We have also pointed out that the scale of the dominant

waves appears to change, being shorter in the main thermocline

than in the deep water. This is also in accord with the WKB ap-

proximation, which predicts that the vertical wave length of an

internal wave should be shorter in regions of greater N. Thus,

it is also possible to use the WKB approximation to normalize wave

lengths in the vertical. This has been done by stretching the

pressure coordinate according to the following differential law:

A= ( Az 11-2N0

where z is the original pressure coordinate, and z is the stretched

pressure coordinate.

Page 38: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

Figure 9 gives z as a function of z, where z is in decibars

and z is in "stretched decibars." Below about 2000 dbar,i(z*) is

approximately a straight line. This, along with the fact that the

high-frequency perturbation kinetic energy is roughly constant with

depth below 2000 dbar, indicates that vertical wave number spectra

of the profiles below 2000 dbar could be approximately calculated

without use of the above normalization and stretching procedures.

But to include the thermocline in these calculations requires the

use of the above procedures.

Figures 10 and 11 show an example of a profile which has had

the above procedures (equations II-1 and II-2) applied to it. Fig-

ures 10 and 11(a) show the velocity profile 219D as it was origi-

nally obtained. Figures 10 and 11(b) show the same profile after

the "geostrophic" shear flow has been removed. Finally, Figures 10

and 11(c) show 219D after the above procedures for stretching and

normalizing the profile have been used. With the exception of

values in the top 100-200 stretched decibars, these profiles pre-

sent a much more uniform appearance than do the profiles of the

east and north components of 219D before stretching and normalizing.

The departure from uniformity in the top 200 stretched decibars

could be explained either by the fact that the WKB approximation

breaks down near the surface, where N is changing rapidly, or by

the fact that the shear of the mean horizontal velocity increases

near the surface. This shear is not taken into account by the above

WKB normalization.

Page 39: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

PRESSURE (STRETCHED DEC/BARS)

I000

oi 2000

3000

q, 4000

5000

6000

Figure 9. Stretched and unstretched verticalpressure coordinates.

Page 40: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

EA T (CM/) EA S T (NCM I/S

) -15 -10 -5 0 5 10 15 20 -10 -5 0 5

400-

,800-

1200-

-1600 -

20001

C

Figure 10. East component of profile 219D as originally obtained (a), after removal of themean profile (b), and after stretching and normalizing (c).

-20 -1

1000

2000

3000

5000

Lu154

(Zi44

a b

Page 41: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

NORTH, (CM/S) NORTH, (NCM/S)

Figure 11. North component of profile 219D as originally obtained (a), after removal of themean profile (b), and after stretching and normalizing (c).

(i)

tj

LU

J

Page 42: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

f.) Spectral decomposition of stretched profiles

The time series of profiles is now in a form (the stretched

and normalized profiles, such as 219D in Figures 10 and 11(c)) in

which the vertical wave number spectrum of horizontal high-frequen-

cy kinetic energy can be calculated. We now describe and present

several different types of spectra which have been obtained.

The first type, shown in Figure 12, is the simple autospec-

trum, or the total energy of the east and north components as a

function of stretched vertical wave number. For this spectrum, a

subset of nine down profiles out of the original 20 was selected.

The only reason for this selection was that some profiles have

better data near the surface than others. The EMVP quite often

does not obtain accurate velocity values for the first 50-70 dbar

below the surface. Since the WKB procedure described above tends

to weight values near the surface more heavily than val.ues in

deeper water, it was important to choose those stretched profiles

which have the best data near the surface. Thus, the best nine

down profiles were chosen for the calculations. The top 50 dbar in

the stretched profiles were also ignored when calculating the spec-

trum. This further reduced the influence of near-surface measure-

ment errors. The 95% confidence limits were calculated using an

assumed 36 degrees of freedom (Jenkins and Watts, 1968). Each

final spectral estimate is formed by averaging the original esti-

mates over four adjacent wave number bands in each profile, and then

averaging over the nine profiles. This would give 72 degrees of

Page 43: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

WAVE LENGTH(SOB)

100

0.01

VER TICA L WA VE NUMBER(C/SDB)

Total spectrum (east and north energy) for 9 down profiles.

LQK

'4

(%)

cN4

102

20

0.0510

Figure 12.

Page 44: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

freedom to each estimate, if east and north velocities are statis-

tically independent. However, Figure 13 indicates that the east

and north velocity components are not independent. Therefore, 36

degrees of freedom appears to be more realistic.

The somewhat unusual units in the spectrum of Figure 13 arise

because of the normalization by N(z). Thus NCM/S means "normalized

cm/sec," and C/SDB means "cycles per stretched decibar." If m and

m' are vertical wave numbers in the unstretched and stretched coor-

dinates, respectively, then the relation between them is

m = ( ) m' . 11-3

Normalized cm/sec are the units obtained when equation II-1 is ap-

plied to an original velocity profile.

The most interesting characteristic of the autospectrum is

that at high wave numbers it has a slope of about ,2.5 on a log-log

plot. At smaller vertical wave numbers this slope decreases. This

spectral shape will be compared in Chapter V with a theoretical

vertical wave number spectrum derived by Garrett and Munk (1972,

1975).

The second type of spectrum which is presented is a form

which treats the horizontal velocity vector (u,v) as a complex vec-

tor, u + iv, (Gonnella, 1972; Mooers, 1973). At any vertical wave

number, m', the Fourier transform of u and v separately gives sinu-

soids for each component. These two sinusoids, taken together,

will form an ellipse in the u, v plane, the position of the velocity

Page 45: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

WAVE LENGTH(SDB)

100

0.01

20

0.05

VER TICAL WA VENUMBER(C/SDB)

Figure 13. Clockwise (CW)down profiles;

and counterclockwise (ACW) spectra for 9In the text, CW is C(m') and ACW is A(m').

103

C)

c

144

X 1(%)

101

10

Page 46: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

vector on the ellipse being a function of z , the pressure in

stretched decibars (increasing downward). Since any ellipse can

be represented as the sum of two complex vectors rotating in op-

posite directions, but with the same frequency, the complex vector

umr + ivm, can be represented by

im'z* -im'tvum + iv, = u e + u e II-4

where m' is always considered to be positive, and u+ and u_ are

complex. The energy spectrum is then given as a decomposition be-

tween the clockwise energy, C(m'), and the counterclockwise energy,

A(m'), defined by:

C(m') = 1/2 u , u > ; II-5

A(m') = 1/2 < u+ u+ > , II-6

where the brackets denote an average over realizations (in this

case, profiles) and possibly over m', the stretched vertical wave

number. The total energy, T(m'), is given by

T(m') = A(m') + C(m'). 11-7

T(m') is plotted in Figure 12, and A(m') and C(m') for the same

nine drops are plotted in Figure 13.

Probably the most striking aspect of Figure 13 is the fact

that clockwise energy is greater than counterclockwise energy over

much of the stretched vertical wave number range. Only in the high

wave number region of the spectrum, where the energy is relatively

Page 47: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

low, do we find points where counterclockwise is greater than

clockwise energy. This means that the waves are elliptically

polarized in the clockwise sense, or in other words, that there is

a distinct tendency for the horizontal current vector to rotate in

a clockwise sense with depth, as seen by an observer looking down

from above. This has important implications for the theory of

internal waves, which will be discussed in the next chapter.

The third type of spectrum that we present here is simply

that obtained by plotting the energy of the east and north compo-

nents separately. This spectrum is given in Figure 14 for the

same nine down drops as in Figure 12. We see that, in contrast to

the difference in Figure 13 between the clockwise and counterclock-

wise spectra, Figure 14 shows that there is almost no difference

between the energy of the east and north velocity components. This

is an indication of isotropy; that is, that in the mean there is as

much energy in the east-west as in the north-south direction.

Finally, we present in this section results of spectral cal-

culations performed on nine up profiles from the same time series.

It often happens that on a single drop (up and down) the down part

will have good near-surface data while the up part will not. For

this reason, the nine best up profiles chosen do not coincide with

the nine best down profiles. These spectra are shown in Figures

15, 16 and 17. Each spectral estimate represents an average over

four adjacent wave number bands and over nine up profiles. The

major reason for performing the same calculations for up and down

Page 48: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

WAVE LENGTH(SDB)

100 20

0.01 0.05

VER T/CA L

Figure 14.

WA VE NUMBER(C/SDB).

Spectra of east and north velocitycomponents for 9 down profiles.

K,

102

I0'

C)

XP

10

Page 49: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

WAVE LENGTH(SDB)

100

0.01

Figure 15.

VER 7TCAL -WA VE NUMBER(C/SDB)

Total spectrum (east and north energy) for 9 up profiles.

LLJ

LU

Q:

czK1

X

C\4

10"

102

20

0.0510

Page 50: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

WAVE LENGlTH(fSDB)

100

0.01

20

0.05

VER 7TCAL WAVE NUMBER(C/SDB)

Figure 16. Clockwise (CW) and counterclockwise (ACW) spectra for9 up profiles.

k

C~j

(%)

K-

102

10'

X0

C\

100

Page 51: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

WAVE LENGTH(SDB)

100 20

0.01 0.05

VER TI CA L WA VE NUMBER(C/SDB)

Figure 17. Spectra of east and north velocitycomponents for 9 up profiles.

103

102

K

LU

(%)I:::l4~J

C

(I10

10

Page 52: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

profiles was to show that there is very little difference between

them (even though all of the same profiles were not used in both

calculations). That the two computations are very similar can be

seen from comparing Figures 12 and 15, 13 and 16, and 14 and 17.

The 95% confidence limits (with 36 degrees of freedom) have

been included for the total vertical wave number spectra (Figures

12 and 15) and for the spectra of east and north velocity compo-

nents (Figures 14 and 17). These confidence limits should be used

with caution, however, since they strictly apply only to a white

noise spectrum. The spectra presented in this section are clearly

"red." Even though the original profiles were pre-whitened (by

taking first differences) before applying the 7ourier transform,

and the above spectra were then produced by re-coloring the spectra

obtained from the first-difference pfofiles, it is likely that the

pre-whitening did not completely convert the profiles into ones

with flat spectra. This means that the actual number of degrees

of freedom could be less than 36. We have also not plotted confi-

dence limits for the clockwise and counterclockwise spectra.

The clockwise and counterclockwise energy estimates depend not only

on the probability distribution of the autospectra for the east and

north velocity components, but also on the probability distribution

of the quadrature spectrum between east and north velocity compo-

nents (Appendix D). Instead of deriving a theoretical estimate for

the probability distribution of the clockwise and counterclockwise

spectra, it is probably more important to point out that, as

LILILYl~__laYIJI__lr^..._C Y-III~LIIII I .~-l._)i- l__ ~^__.~--~.--ri--il1liL

Page 53: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

indicated below, clockwise and counterclockwise spectra calculated

from four-drop groups in the time series always showed that clock-

wise energy was greater than counterclockwise energy over most of

the stretched vertical wave number range. Furthermore, while av-

eraging the spectra of two four-drop groups together gave clockwise

and counterclockwise spectra very similar to those of Figures 13

or 16, averaging three four-drop groups together did not appreciably

change the spectra. This indicates that nine profiles are enough

to give stable estimates of the clockwise and counterclockwise

spectra.

Before the spectrum of Figure 12 was calculated, five pre-

liminary autospectra were obtained by using four-drop groups in

the series of 20 down profiles. That is, drops 219D-223D were used

to obtain a spectrum, then drops 224D-228D, and so on. All four-

drop groups showed a dominance of clockwise over counterclockwise

energy. Each of the five preliminary spectra also showed a tend-

ency for the energy to peak at vertical wave lengths of 100-130

sdb. Drops 230D-233D, in particular, showed a strong energy peak

there. However, in some cases (for example, 219D-223D) this peak

did not show up as clearly. The fact that the energy spectra did

appear to have some dependence on time indicates that the spectra

calculated from the nine best profiles out of the 20 will tend to

smooth out peaks in the energy spectra.

On the other hand, it is clear that energy contributed at

vertical scales of 100-130 sdb is an important part of the stretched

I~IUlrl_ i L~*~l_~__l~l~ _ ~I___ _ 1.111~---

Page 54: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

profiles. From Figure 3 we see that the most apparent visual var-

iability in this profile occurs in the above range of vertical

scales. (In the original pressure coordinate, the above scales

correspond to 100-150 dbar in the main thermocline, and reach

1000 dbar in the deep water.) A comparison of drops 219D and 2210

in the stretched vertical coordinate also shows that much of the

"mirror-imaging" that occurs between these profiles is caused by

waves with the above vertical length scales.

g.) Paired profiles

During the course of the time series experiment, several

simultaneous paired profiles were made, using two EMVP's. These

were done at separations of from 100 meters to about 15 kilometers.

Figure 18 shows two profiles which were made simultaneously at a

separation of 4.8 kilometers. Profile 235U was made to the south

of profile 236U. It can be seen from this figure that many velocity

features of 235U can be traced visually to corresponding features

in profile 236U, although there is quite often a depth change be-

tween the more energetic features of profiles 235U and 236U. This

indicates that: a.) many of the features observed in the two

series (at least those with wave lengths greater than about 100

dbar) have a large ratio of horizontal to vertical length scales;

and b.) the energetic features in the two profiles are not hori-

zontal, but are slightly inclined. We will return later to a con-

sideration of the inclination of velocity features between 235U and

Page 55: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

VELOCITY PROFILESSIMULTANEOUS

235U & 236UDROPS 4.8 KM APART

EAS7 (CM/S) NORTH, -. (CM/S)

()

Figure 18. Simultaneous paired drops, 235U and 236U, separatedhorizontally by 4.8 km.

Page 56: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

56

236U. We point out here, however, that waves with large ratios of

horizontal to vertical wave lengths should, according to the re-

sults of Appendix B, be measured with small error by the EMVP.

Page 57: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

Chapter III Vertical Polarization and the Vertical Propagation of

Internal-inertial Waves

a.) Summary of data

Before going further, it will be useful to summarize the

results of the previous chapter. Figure 3 indicates that high-

frequency energy in the profiles is dominated by waves near the

local inertial frequency. There is probably also energy contri-

buted by diurnal and semidiurnal tides. Figures 6 and 7 show that

the velocity structures associated with these inertial waves

clearly move upward with time at the time series location. These

two figures also show that the east and north components of hori-

zontal velocity are usually not in-phase with depth, but that the

east component lags behind the north component. This indicates

that these near-inertial waves are polarized with depth*, a fact

which is born out by the clockwise and counterclockwise spectra of

Figures 13 and 16. These spectra show that the waves are ellipti-

cally polarized with depth, and they show that this polarization

is predominantly clockwise (the clockwise energy is greater than

the counterclockwise energy over most of the observed wave number

band). We note that the observed lag in Figures 6 and 7 between

the east and north components also indicates that these waves are

polarized clockwise with increasing depth. Finally, Figures 14 and

17 indicate that there is very little difference between the energy

of the east component and that of the north component. We would now

Page 58: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

like to see how the above observations may be related to internal

wave theory.

b.) The vertical spatial behavior of horizontal internal wavevelocity components

Expressions for the east and north, or u and v velocity

components of a single propagating internal wave are given in the

WKB approximation by:

S= ( ak i ifl) x A (N2 () _ 2) 1/4

v ( + ikf T wl) ) (2 _ f2)l/ 2 (k2 + 12)1/2

k2 + 121/2 1/2+ i(- (N2() - 2) dzS 2 f2 (N i(kx + ly - wt)

xe x eIII-1

where A is the wave amplitude, and tie coordinate system is chosen

so that x is positive eastward, y is positive northward and z is

positive upward. The coordinates 2 and 2* will be reserved for the

vertical pressure coordinate (increasing downward) in decibars and

stretched decibars, respectively. (The reader may refer to Chapter

V for the derivation of these formulas.) In order to simplify

these expressions, we rotate the coordinate system through an angle

a as shown in Figure 19. In the new coordinate system:

0' = 1 cosa + 9 sina ;

v' = - sina + v cosa ;

k' = k cosa + 1 sina ;

l' = -k sina + i cosa = 0 , 111-2

Page 59: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

44,9

t

v I /"0 ,'%

u KX

Figure 19. Horizontal coordinate rotation (see equations 111-2).

-h

.4-,

I

Figure 20. Direction of rotation of the horizontal velocity vector of aninternal wave for positive and negative vertical wave numbers.The curved arrow shows the direction in which v rotates withincreasing z.

~LI~I iliuuruu~--l*i~-*~r+-p- ~~...~_li--r-------cI~ -- ~P--L~- ---~Y__

ii'

IKr

Page 60: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

where

COSo = kik 2 + 12

and

sina =/ k2 + 12

In the new coordinate system, we have

u' = 1 xA (N(z) 2 /4

S f 112(z) /2)+ i- (W2 _ f21/2

+ i f2) 1 2 f(N2(z) _W2) 12

z i(k'x' - at)x e

k 2 1/2k' 2 2

2 /2- 2) dz Z m(z) z , III-5

where m(z), the local vertical wave number, is given approximately

by

m(z) 2= k 22)1/2 (N2(z) - 2)/2

Note that m(z) is a positive quantity. The choice of signs in the

expressions for horizontal velocity determines whether the wave is

111-3

x e

We let

SIII-4

111-6

Page 61: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

propagating upward or downward. Then

(N2 2) 1/4 i(k'x' + mz - wt)u' = l (N2(z) -W )9' =+ --f (2 - f2)/2 -7

- {O

The horizontal component of the wave number vector, k', is assumed

to be in the direction of horizontal phase propagation, so w > 0.

We note in passing that i' and ' are proportional to N1/ 2, while

m(z) is proportional to N. This confirms the connection made in

Chapter II, without proof, between the WKB approximation and the

procedures used for stretching and normalizing the velocity pro-

files. Taking real parts of 5' and 9':

Re (5') = u' = + cos (k'x' + mz - t) xA (2(z) - 2) 1/4

Re (9') = v' = -f sin (k'x' + mz - wt)S (wZ2 f 2)/ 2((W

III-8

The upper signs are chosen if m is directed upward, and the lower

signs are chosen if m is directed downward. A is assumed to be

real. To observe the behavior of u' and v' as z varies, we set

x' = t = 0. Then

1/4u' = F cos (+ mz) xA (N2(z) - ) 1/4

v' = - sin (( mz) (2 f2)/ 2 III-9

The two cases for the choice of upper or lower signs are shown in

Page 62: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

K

*Cg

C9

KFigure 21. Propagation diagram for internal waves. Cg is the group

velocity vector, k is the wave number vector, and v is avector representing water velocity.

_II__IWQYJIP_____III- Il~i--X- - - ~

Page 63: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

4. 4 2 1/4Figure 20. We note that ' V'/(N2(z) - . The horizontal

4.vector, v', is plotted in Figure 20 instead of v' in order to re-

move variations in the horizontal velocity vector magnitude brought

about by changes in N(z). The ellipses represent hodographs of the

horizontal velocity vector as z varies. As w -+ f, the wave becomes

circularly polarized. We note that although we have been writing

simply A for the wave amplitude in the above expressions, the wave

amplitude is more generally a function of vertical wave number,

frequency, and a, the direction of horizontal phase propagation;

that is, A = A(m, w, a). In the + m case, the horizontal velocity

vector rotates counterclockwise with increasing z, hile in the - m

case, the sense if rotation is the opposite. The result is that

there is a unique relation between the sign of m and the sense of

polarization in the internal wave hodographs of Figure 20.

We can now relate the sign of m(z) to the direction of the

vertical'group velocity component for internal waves. Figure 21

shows a propagation diagram for internal waves. The vector t is

the wave number vector for the wave and ig is the group velocity

vector (or the direction of energy propagation).

For simplicity, only two wave number vectors have been

shown, one pointing toward the sea surface (toward positive z) and

the other pointing away from the sea surface: the dispersion rela-

tion gives

W2 = N2 sin2 (0) + f 2 COS 2 (0), III-10

I_ _

Page 64: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

644.

where 0 is the angle that the wave number vector, k, makes with

the z axis. For w close to f, 0 is small, and t is almost verti-

cal. For constant N, i is also independent of depth. On the other

hand, if N varies with depth, then e and the vertical component, m,

of k also change with depth. If the wave propagation may be approx-

imated by a WKB expansion, however, equations III-10 and III-11 re-

main true locally, except that m and N are now functions of z. The

magnitude of the group velocity is given by:

Cg 1 a : (N2 _ f2) sin-0 cos 0

I'l a IIl [N2sin20 + f2cos2 O]1/2 III-11

The group velocity vector is perpendicular to i and points away

from the z axis if N > f, which it was everywhere in the water

column during the time series experiment. The vector v in Figure

21 is the water velocity due to the wave. Since the water is as-

sumed to be incompressible, i v = 0 always. There is no depen-

dence on horizontal direction in the above formulas, so the t and

tg vectors can be rotated about the z axis without change of fre-

quency or magnitude of tg. We have also drawn two helices in

Figure 21. These helices represent the path that the tip of the

velocity vector, v, makes as distance from the origin is increased

along k.

Looking down on this figure from above the z axis, it is

readily seen that the sense of polarization of the wave (that is,

the direction in which v rotates with increasing depth) changes,

Page 65: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

654.

depending on whether k is directed toward or away from the sea sur-

face. For a vector k directed toward the 5ea surface, the wave is

polarized clockwise with increasing depth, while for a k vector

directed away from the sea surface, the wave is polarized counter-

clockwise with increasing depth. This is simply a restatement of

the result expressed by equations III-8 and III-9.

We point out that polarization can be inferred easily from

the behavior of internal waves in time. We assume we are observing

at a fixed point within the ocean in the Northern Hemisphere. Then

it is known that a velocity vector due to an internal wave propa-

gating through that point will rotate in a clockwise sense with

time, due to the influence of the Coriolis acceleration. Since

the internal wave is assumed to be propagating along the direction

of the wave number vector, t, of Figure 21 (either toward or away

from the sea surface), the waves have to be polarized as shown in-.

the figure in order to give a velocity vector, v, which rotates

clockwise in time at a given point in space.

Now, Figure 21 also shows that if t is pointed toward the

sea surface, Cg is pointed away from it, and vice versa. There-

fore, if either the sense of wave polarization or the direction of

the vertical phase propagation (the sign of m) can be determined,

then the direction of the vertical component of the group velocity

vector, and thus the direction of vertical energy propagation, can

be found (Leaman and Sanford, 1975).

Page 66: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

The data presented in Chapter II allows us to determine both4.

the direction of the vertical component of k (at least for the

dominant waves) and the sense of wave polarization in the vertical.

As already pointed out, the dominant waves in Figures 6 and 7 al-

most always appear to be propagating upward through the upper 2.5

kilometers of the water column. This indicates that the direction

of phase propagation, or i, for these waves is pointed toward the

sea surface. If this is true, the above discussion shows that the

waves should be polarized predominantly clockwise with increasing

depth. Both the clockwise and counterclockwise spectra of Figures

13 and 16, and the fact that the east and north velocity components

are out-of-phase in depth, with north leading, in Figures 6 and 7,

show that the waves are in fact dominated by clockwise polarization.

The influence of the mean flow on the observed waves has so

far been ignored. The depth dependent part of the mean flow pro-

file has already been presented (Figures 4 and 5). Another depth

independent component of the low-frequency profile may also exist,

but its magnitude cannot be evaluated from EMVP data (Sanford,

1971). In order to evaluate a possible depth independent part of

the mean profile, we have calculated the average horizontal cur-

rent at the central mooring as measured by a current meter at

1513 dbar (1496 meters). The average horizontal velocity compo-

nents over the interval of the profiler time series, as determined

from the current meter record, are + 1.6 cm/sec to the west and

+ 2.5 cm/sec to the north. The magnitudes of the horizontal current

- - , - 0, - , , 6 .6 'dM --- - - , __ __ -

Page 67: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

components at 1513 dbar in Figures 4 and 5 are ~ 1.5 cm/sec to the

west and ~ 1.5 cm/sec to the north. This indicates that there is

very little shift in the east component, while - + 1 cm/sec should

be added to the north profile to make the average north flow ob-

served by the current meter and the profiler agree.

Since the wave number vector, t, is not exactly vertical

for waves propagating in the vertical, we expect that the advection

of waves by the mean flow will cause an apparent vertical propaga-

tion. The amount and sign of this apparent propagation (upward or

downward) depends on the direction of horizontal wave propagation

relative to the mean flow. Figures 4 and 5 show that the main part

of the mean flow is oriented north-south. Thus, for example, if

the waves were propagating east-west, there would be little appar-

ent vertical propagation due to the advection of waves past the

observing point. On the other hand, if most of the waves were

propagating northward (and upward), the advection would cause an

apparent downward phase propagation to be added to the true upward

phase propagation of the wave above 1500 dbar. In the absence of

adequate information on the direction of horizontal phase propaga-

tion it is difficult to estimate the influence of the mean flow.

There is some theoretical support for assuming that the near-iner-

tial waves are propagating northward (Munk and Phillips, 1968).

If this is, in fact, true, Figures 6 and 7 show that k must be di-

rected toward the sea surface: that is, the actual upward phase

propagation of the waves is enough to overcome any apparent

Page 68: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

downward propagation caused by advection. A strong indication

that the observed waves are propagating upward can be found by ex-

amining Figures 4 and 5. It is known from tracking neutrally

buoyant floats that the mean, or "low-frequency" flow is small at

depths of around 1500 dbar (Voorhis, 1974). Figures 4 and 5 also

show that, taking into account the shift in the north profile re-

quired to make the mean flow as observed by the profiler and cur-

rent meter equal at 1513 dbar, the mean flow in the depth range

1400-1500 dbar is small (< 3 cm/sec). Therefore, there should be

little advection of waves at this depth. However, Figures 6 and 7

show that the waves are moving upward at this depth. In fact, the

rate of upward motion is greater at 1500 dbar than it is at lesser

depths. Vle note that the sense of polarization is not affected by

the advection of waves past the observation point. Therefore, for

the purpose of determining the direction of vertical wave energy

flux, observation of polarization appears to be more useful than

observation of the apparent vertical phase propagation.

The behavior of wave polarization and the direction of

phase propagation both appear to indicate that the vertical energy

flux is directed downward, away from the sea surface. It is nec-

essary, however, to discuss just what the partition of energy be-

tween clockwise and counterclockwise means for these waves. In

particular, we have limited the above discussion to a single wave.

It is more appropriate to relate the partition of clockwise and

L -LLillllslC -~d~~^---CI" IPt-~ -C-~~^~ I

Page 69: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

69

counterclockwise energy to the energy carried by upward and downward

propagating waves. This will be done in Chapter IV.

It is known from internal wave theory (see, for example,

Garrett and ,unk, 1972) that internal waves with frequencies near

the inertial frequency are isotropic in time; that is, the magni-

tude of the frequency spectrum of a near-inertial wave horizontal

velocity component will be independent of the direction of that

velocity component. The same isotropy will be true for the verti-

cal wave number spectra of near-inertial waves, provided only that

east and north spectra for a sufficient number of profiles are

averaged together.

To see why, this is true, we consider a single internal wave

propagating energy downward. The u' and v' velocity components are

given by equations 111-8:

uL = - cos (k'x' + mz - wt) (N2 - )1/4

V1 - sin (k'x' + mz - wt)~ (2 - f2) 11 2 111-12

where A is the wave amplitude. -We-point-ot- that the hodograph in

z for this wave is given by:

u' 2 2 (N2 - W2)1/2

1 + a A2(m,,)( )2 ( 2 - ) 11-13

At any time, the ratio of semi-major to semi-minor axis is - , andf

as w - f, the wave becomes circularly polarized. If we assume the

presence of a second wave propagating upward with the same vertical

Page 70: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

wave number magnitude, m, and the same frequency, w, but with dif-

ferent amplitude, B, then the velocity components for this wave are

given by:

u= cos (k'x' - mz - wt) x1/4u x (N2 .)2)

vu f sin (k'x' - mz - t) (2 f2) 1/2 I-14W - 111-14

We assume A and B to be real now, although this is not necessary.

We also assume we are observing at the coordinate origin, so

x' = 0. The result of adding the two waves together gives:

1/4S (N2 )2

UT= (/2 [(B - A) cos(mz)cos(wt) - (A + B) sin(mz)sin(wt)],

(W2 - f2)

1/4

v' = (N - ) [(A - B) cos(mz)sin(wt) - (A + B) sin(mz)cos(wt)].(W2 - f2) /2

If-15

Weisee that, for a fixed z, the magnitude of u' averaged

over a cycle in time, becomes equal to the average of v' as w - f

independently of the magnitudes of A and B. Thus, spectra of u'

and v' in frequency may be expected to show the same energy when

the frequency is close to f.

The same is not true, however, if vertical wave number spec-

tra of u. and v are considered, as can be seen by fixing t in the

above expressions. If A or B is zero, the vertical wave number

spectra for u' and v' will be approximately equal for w close to f.-r

IIIIIL_-LII~I~_.~ ~_ -~.--.I~ ~

Page 71: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

However, if A = B (which would occur if standing waves existed in

the vertical), energy in the vertical wave number spectra of u' and

v.' will depend on t, the time in the wave cycle at which the spectra

were obtained. For example, in this case, at t = 0, the spectrum

of v" will contain all the energy at that particular vertical wave

number, m, and the spectrum of u' would, in theory, be zero.

However, if vertical wave number spectra of u' and v' are_r _

taken at enough different times, t, in the wave cycle, we again ap-

proach isotropy for w -+ f. This is true because, at a given m, the

energy in u' and v' will be the same if averaged over time. Ther _T

average vertical wave number spectra of Figures 14 and 17 can be

considered as the sum of individual wave numbe- spectra taken at

numerous different times, t, in a wave cycle. Then the fact that

much of the energy in the profiles apears to be contributed by

waves with frequencies close to f explains why there is an equi-

partition of energy between the east and north components.

c.) Polarization in the deep water

As mentioned previously, we can approximately calculate deep

water spectra (below 2500 dbar) without stretching and normalizing

the velocity profiles. This is possible because N does not change

much in the deep water. Calculating the clockwise and counter-

clockwise spectra over the depth range 2500-5500 dbar for several

profiles indicates that dominance of clockwise energy over counter-

clockwise energy exists in the bottom half of the water column, at

least for vertical wave lengths greater than about 100 dbar. This

-i--L~1' ~--sur~ u*rp*_^-,r-r~i.~.... -~-I*--.--- ~ ̂-^. ~----- i ---*- e~-rr.

Page 72: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

is important because it indicates that, even in deep water, the

net flux of energy by these near-inertial waves is downward.

Pollard (1970) indicated in a theoretical study that it would be

difficult for near-inertial waves to propagate down into the deep

water with the observed intensity. The result of the above cal-

culations indicates that these waves are, nevertheless, coming down

from regions farther up in the water column. The Pollard model

assumes that inertial waves are generated at the surface by an im-

pulsive force that acts over some time and space interval. In

Chapter V we will examine the behavior of the vertical energy flux

of near-inertial waves for an assumed form of the internal wave

energy spectrum in frequency and vertical wave number space.

d.) Comparison with meteorological observations

As was pointed out in the introduction, meteorologists have

been conducting experiments to obtain vertical profiles of atmos-

pheric winds for many years. Their results may be compared with

results obtained by the EtMVP in the ocean.

Various authors have pointed out that the existence of in-

ternal waves in the atmosphere has not been proved. For example,

Justus and Woodrum (1973) felt it necessary to state that: "There

has been no unambiguous resolution of gravity waves in the upper

atmosphere (i.e. simultaneous observations of amplitudes, phases,

and frequency sufficient to verify propagation according to the

theoretical dispersion equation)." They go on to state that there

IIU____CIYI___I___~_i~-C~ LIII_ i I~l _. ~X ~~-~ I~IUYL~

Page 73: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

is, however, strong circumstantial evidence for the occurrence of

internal waves. Some of that evidence will now be presented.

In 1966 radar tracking of radar-reflective balloons was used

at Cape Canaveral to obtain profiles of atmospheric winds from

0 kilometers to about 16 kilometers (Endlich, et. al., 1969). Two

series of profiles were made, each series lasting between one and

two days. The method allows one to sample rather rapidly in the

vertical (one point each 25 meters) but, unfortunately, it can

only determine the wind speed at each point, not its direction.

The authors point out that there are strong velocity features in

the July 4-5, 1966, experiment which progressively move downward

with time. The profiles are complicated, however, and no definite

time or space scales stand out (although it is known that internal

waves can exist in the atmosphere at these altitudes).

Although apparent high-frequency features in the atmosphere

have been observed to change altitude, there appear to have been

few reports of any polarization or directional structure associated

with these features. Endlich could only measure speed as a func-

tion of altitude and therefore could not observe polarization.

Hines (1966) has reported polarization of atmospheric winds that is

apparently related to the diurnal atmospheric tide. His data, ob-

tained by tracking sodium vapor trails at heights of 90-130 kilo-

meters, indicates that winds with periods appropriate to the diurnal

tide rotate with altitude in a given rocket profile. If an observer

is imagined to be standing above the atmosphere and looking down,

II~L1 YI_~ I_~L _ CJI^^___I__IY__I_~ PYY-

Page 74: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

then the waves seen by Hines have horizontal velocity vectors that

rotate counterclockwise with increasing "depth," or atmospheric

pressure. This is the opposite of the case observed in the ocean

by the EMVP over the smooth topography, in which the horizontal

current vector rotates clockwise with increasing depth. As Hines

points out, the atmospheric observations indicate that the wave

energy of these oscillations is propagating from lower altitudes up

to the 90-130 kilometers range. This appears to be one of the few

cases where wave polarization has been observed in the atmosphere

(C. 0. Hines, 1975, personal communication).

One good reason for a lack of elliptical or circular polar-

ization in the motions interpreted as being internal waves in the

atmosphere is the fact that these motions quite often have periods

much shorter than the local inertial period. Hines (1960) points

out that atmospheric internal waves might be expected to be most

energetic at periods of several hours. For waves of this period,

the earth's rotation is not a major influence, and a single atmos-

pheric internal wave with a three hour period would be almost

linearly polarized in the vertical.

Hines also notes that some observations in the atmosphere

show structures which progressively move downward in time. Using

the same arguments as those put forth in describing Figure 21, he

points out that because the directions of vertical phase propaga-

tion and energy flux are opposite for internal waves, these down-

ward moving structures in the upper atmosphere could be interpreted

I~ I^IC~Llli-O-IWI I(C- --U1 X~I-l~~llll^n-^~ YTrIIYYIII XI-~l~-l-- ~~ --LI_-lllll~l llll

Page 75: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

as internal waves carrying energy upward, away from the lower at-

mosphere and the surface of the earth. (The sources for atmospheric

internal waves are thought by many meteorologists to be located in

the lower atmosphere or at the earth's surface.)

Finally, there is some evidence from fresh water lakes that

appears to indicate upward phase propagation of internal waves.

Lazier (1973) obtained a series of temperature profiles in Lake

Bala, Wales, over a time interval of about 12 hours. Profiles were

repeated at approximately 11 minute intervals. In many cases,

features in the temperature profile were observed to move progres-

sively upward in time. Lazier attributes this upward motion of

temperature featt.res to the presence of propagating internal waves.

If the energy source for these waves was located at the lake sur-

face, they would have to propagate their energy downward. Then,

as in Figure 21, the wave number vectors of the waves wbuld be

directed toward the lake surface, and the wave phase would move

upward in time. Since Lazier only measured temperature profiles,

however, no data on wave velocity polarization were obtained.

Page 76: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

Chapter IV The Use of the Ratio of Clockwise to Counterclockwise

Energy as a Reflection Coefficient. Reflection of Internal Waves

From a Smooth Bottom

a.) Calculation of a reflection coefficient for the observed waves

We pointed out in Chapter III that the dominance of clock-

wise over counterclockwise vertical wave number spectra for hori-

zontal kinetic energy appears to indicate that the net vertical

energy flux for these waves is downward. However, a further step

must be taken if we are to relate the partition between clockwise

and counterclockwise energy to the energy in upward'and downward

propagating waves. This step is to assume that the observed waves

have frequencies close enough to f so that any single wave should

be almost circularly polarized in the vertical (see equation 111-13).

The sense of polarization is determined by whether the wave energy

flux is downward (clockwise polarization) or upward (counterclock-

wise polarization). With this assumption, the horizontal kinetic

energy of the downward propagating waves is approximately equal to

C(m'), the clockwise energy in the vertical wave number spectrum,

while the horizontal kinetic energy of the upward propagating waves

is approximately equal to A(m'), the counterclockwise energy in the

same spectrum. (Strictly speaking, A(m') and C(m') should be multi-

plied by the Brunt-Vaisila frequency to get the true energy level

(see Chapter V). However, since we are finally interested only in

~YY~ ?~l_~~ _il~LI~- ~------C

Page 77: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

the ratio of clockwise to counterclockwise energy, this step has

been ignored.)

We have assumed in the previous paragraph that inertial

waves are the primary cause of high-frequency energy in the pro-

files. However, frequency spectra of horizontal velocity obtained

at the central mooring (M. Briscoe, 1975, personal communication)

indicate that the diurnal and semidiurnal tides also contribute

to the internal wave frequency range.(see Figure 29). Because of

the relatively short length of the time series, it is difficult to

resolve semidiurnal tidal and inertial oscillations in frequency,

and it is impossible to resolve diurnal tidal and inertial oscilla-

tions. On the other hand, we can make inferences concerning the

probable influence of the tides on the vertical wave number spectra

computed in Chapter II. First, that part of the tidal energy which

is caused by the surface or barotropic tide, since it is depth-

independent, is not sensed by the EMVP. Second, work of other in-

vestigators (Hendry, 1974) appears to show that most of the tidal

energy, at least for the semidiurnal tide, is concentrated in the

first two or three baroclinic modes (along with the surface tide).

In terms of the vertical wave number spectra of Chapter II, this

means that any tidal contributions to the spectra are located in

the lowest and, possibly, the second-lowest wave number estimates.

We may then infer that above the first one or two vertical wave

number estimates, the primary contributions to the spectra come

from inertial-internal waves. There is also direct evidence from

iq~

Page 78: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

the profiles to indicate that most of the semidiurnal tidal energy

is contributed by waves with small vertical wave numbers. This

evidence will be presented in Chapter V.

We now assume that the observed waves are generated at or

near the surface and propagate their energy downward. Upon reach-

ing the bottom, some fraction of their energy is reflected upward

toward the surface, while the remainder of the energy is lost in a

frictional boundary layer at the bottom. There may also be some

energy loss as the waves propagate through the interior. We also

assume that the smooth bottom does not act as an energy source for

these waves (that is, a boundary layer at the bottom will remove

energy from the incoming waves, but other processes do not act to

put energy into waves that are reflected from the bottom). The

theoretical model presented later in this chapter will hopefully

indicate how reasonable the above assumptions are.

If we accept these assumptions, then the ratio A(m')/C(m')

can be thought of as a reflection coefficient for these waves.

This ratio is shown in Figure 22, using the spectra of Figures 13

and 16. Since the bottom, by assumption, cannot act to put energy

into the waves, a good test of the above model is that A(m')/C(m')

should always be less than one. An examination of Figure 22 shows

that this is substantially true, although at high wave numbers (or

low energy), the ratio A(m')/C(m') behaves erratically. This er-

ratic behavior at low energy levels may be an indication that the

signal-to-noise ratio of the EMVP is becoming small. A low

YI-rr~-wr~ ru~i~ ~--- I-~--~-- ~urr~--c- -- _ n._rri- -~-- --- --~li

Page 79: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

I. 2 3 4S / 2m' sdb (x / O )

Figure 22. Reflection coefficient A(m')/C(m').

Page 80: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

signal-to-noise ratio will cause the east and north velocity compo-

nents to lose correlation (or coherence). Therefore, the quadra-

ture spectrum of the east and north components will decrease.

Since this quadrature spectrum (in stretched vertical wave number)

is proportional to the difference between A(m') and C(m') (see

Appendix D), the clockwise and counterclockwise spectra will tend

to become equal as the signal-to-noise ratio falls. Also shown

in Figure 22 is the cumulative energy of the stretched vertical

wave number spectrum, expressed as percent of total energy. It can

be seen that about 97% of the total energy is contained in waves

whose ratios of A(m')/C(m') are less than one.

It is important to emphasize that the above interpretation

of the reflection coefficient depends strongly on the assumption

that most of the waves observed in the profiles have frequencies

close to the inertial frequency. The short length of the time

series does not permit good frequency resolution. We have already

pointed out that the lowest wave number estimates appear to be

heavily influenced by the semidiurnal tides.(see Chapter V). At

larger vertical wave numbers it is also likely that waves with

frequencies appreciably greater than f are contributing to the pro-

files. We know that a wave with w greater than f is not circularly,

but elliptically, polarized, with a ratio of semi-major to semi-minor

axes of w/f. The presence of such waves in the profiles would pre-

sumably cause the ratio A(m')/C(m') to increase (the addition of a

circularly polarized wave to one which is strongly elliptically

I~--.

Page 81: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

81

polarized results in an elliptically polarized wave). This argument

indicates that A(m')/C(m') is probably an ever-estimate of the re-

flection coefficieft for the near-inertial waves. The reader may

refer to Appendix D for a more complete discussion of the influence

of higher frequency waves on the clockwise and counterclockwise

spectra.

b.) A model for the reflection of internal waves from a smoothbottom

A theoretical model for internal wave reflection can be

developed that supports some of the assumptions made in the pre-

vious section. This model is a variation of one studied by Phil-

lips (1963) that analyzes the reflection of purely inertial waves

from a flat boundary. The present model includes buoyancy effects

but otherwise gives much the same results as Phillips' model.

We start with the usual equations for internal waves, with

buoyancy and rotation:

2u- fv = - P + vV2uat ax

v + fu = - 3R + VIvat ay

aw a- + b + vV2Wat az

l + N2w = 0at

~x +y az . IV-)ax ay + z . IV-l

~II_ I~ j~_jl_______________U__C _I

Page 82: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

The reader can refer to Appendix B for definitions of the

'quantities in the above equations. The only additional terms in

the above equations are the frictional terms vV2u, vV2v and vV2w,

where v is the kinematic viscosity coefficient. The Brunt-V'isala

frequency, N, is assumed to be constant. This is not a serious

assumption, since near the bottom where we expect the model to ap-

ply, N is fairly constant with depth.

If all variables are independent of the y coordinate, a

stream function, 9, can be introduced that satisfies continuity:

u = . k- and w = - xaz x. IV-2

If Jp is substituted into equations IV-1, these equations can be

reduced to two equations in p and v:

-V2 a V2) 2+ N f2~ 3 f = O;at at ax ataz

f -_2 + [_-- - V2 ] v = 0 . IV-3az at

The coordinate origin is assumed to be located at the bot-

tom, so that z = 0 there, and the coordinates form a right hand

system. Then equations IV-3 must satisfy the boundary conditions

u(z = 0) = v(z = 0) = w(z = 0) = 0 or:

S= 4~4= v = 0 at z = 0. IV-4az

PII~ I_

Page 83: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

We look for solutions of the form:

S= A 6kpz ei(kx-wt)

kiz i(kx-wt)v Ve e . IV-5

If these expressions for i and v are substituted in equations IV-3,

there results a sixth order equation in i:

[R2 ] P6 + [2iw'R - 3R2] P4

+[1 - w"2 - 4iw'R + 3R 2 + i N- R] 02

fw

+[ 2 2 + 2i'R - R2 - i NR] = 0 , IV-6f' f

where w' -and R = -. If we assume that R = 0, the charac-f f

teristic equation for P becomes:

2 N2 - 22 _ f2

or j , = + (N2 - f2) 1/2 IV-71,2 2 - f2

These two roots give the z-dependence that would be expected from

two internal waves propagating without friction.

If we now assume R << 1, we can factor out the above two

roots to get approximate expressions for the remaining four:

~-PI~LY- F-~* SL~,~---_ L X -~II ̂ YI_

Page 84: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

2 = - +

R -

(12) - (1N2 - 4' - 2m N2 - 321/2

R2 R wf 2-. _f2 IV-8

If we assume that the - 4 term dominates under the radical, thenR2

the expression for the four remaining roots becomes;

P2 = L (-u' + 1) IV-9

Noting that w is assumed positive, we find that the two boundary

layer roots which give solutions that decay into the interior are

S , i1 -1 (W' r 1)1/2s, 2R IV-10

The four roots of 1 that have been obtained for the two propagating

internal waves and the two boundary layer solutions are then:

1 2 _ f2

- (N2 - 2 1/2.

2 W2 f2

1/2

1/2and ('l)

IV-I1

Page 85: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

We note that the thickness of the boundary layer is determined by

the quantity (w' - 1) 1/ 2/iR, for w' positive and close to one. As

e' + 1, or as w + f, the boundary layer penetrates more and more

deeply into the interior.

The expressions for v and i are now:

kkz k2Z k 3z kz i(kx - wt)i= [A e +Ae + Ae + Ae ] e

1 2 3 4

kiz ki 2z ki z ki z i(kx - wt)v =[Ve + Ve +Ve 3 +Ve J e IV-12

where p, through P4 are given above. The A's and V's must be deter-

mined from the bundary conditions at the bottom (equations IV-4).

The terms with V1 and A, in the expressions for 9 and v give the

amplitude of the incoming wave, while the terms involving V2 and A2

give the amplitude of the reflected wave. The average horizontal

kinetic energy for the incoming wave is given by:

Ei = 12 (1 + 2) IV-13

The average horizontal kinetic energy for the reflected wave is:

Eo = IV1 2 (1 + 2) I IV-14

where

A=l R ' I 1 1

(i-l)(m-1)1 / 2 2(i-1) ( -1)/ 2 +1)1/2

IV-15

___UEM~ ____Xn_____h_______~~__l--rx~~_-

Page 86: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

and

B = w + i+ (

/2(i-1)

S+ 1

W-1)1/2 (+)1/2

IV-16

and

S= (N2 2

1/2

N2 - m2

Using Ei and E , we can form a

zontal kinetic energy:

E /~ (N2 - W2)/2

Ei (m'2-1)1/2

reflection coefficient for the hori-

+ 1 W' -1+ + O(R).

(-1)1/2 (+l)1/2IV-18 IV-18

This reflection coefficient has several characteristics that

are relevant to the present data. First, although for dimensionless

frequencies w' much greater than one, E /E i ~ 1, as w' - 1 this re-

flection coefficient decreases rapidly (Eo/E i does not limit to zero

as w + f due to the approximations made in the above derivation).

This shows that energy in waves near the inertial frequency may be

greatly reduced as these waves reflect off the bottom. Most of the

energy loss suffered by these waves takes place in the bottom fric-

tional boundary layer (a boundary layer, we recall, that grows in

thickness as w + f). For frequencies close to f, E /Ei is very

IV-17

Page 87: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

sensitive to w. Since we cannot resolve w as closely as would be

required to test the above theory, we can only say that the ob-

served "reflection coefficients" (Figure 22) and the possible

boundary layer thickness (described below) can be accounted for if

the frequencies of the observed waves are close enough to f.

Figure 8 appears to indicate that, from 2000 dbar to about

5300 dbar, at least, the average perturbation horizontal kinetic

energy follows closely the average profile of N. This would seem

to imply that the waves are not undergoing strong frictional losses

as they propagate through the water column, and therefore that most

of the energy loss from these waves takes place as they reflect off

the bottom.

There is also a suggestion in Figure 8 that the mean per-

turbation kinetic energy decreases within about 100-150 dbar above

the bottom. This may indicate the presence of a frictional boundary

layer at the bottom. In theory, the horizontal kinetic energy must

go to zero at the bottom. However, the EMVP only measures horizon-

tal velocity to about 10 dbar above the bottom. Thus, it probably

would not see that part of the frictional boundary layer nearest

the bottom, where the greatest energy loss takes place. The error

estimate (Chapter II) calculated for the mean perturbation kinetic

energy profile of Figure 8 indicates that the decrease in energy

near the bottom is significant. We again must emphasize, however,

that this error estimate only takes into account the influence of

instrumental noise.

--lll^--Yn~)l~-DI llP ~L1__~I___~C

Page 88: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

Figure 8 also shows an apparent bulge in the mean pertur-

bation horizontal kinetic energy at about 200 dbar above the bot-

tom. This bulge remains even if the kinetic energy profile is nor-

malized by the mean Brunt-V'is l profile (N also increases as the

bottom is approached). The theory presented above predicts only

that the mean perturbation horizontal kinetic energy should increase

smoothly from zero at the bottom to values appropriate to the in-

coming and outgoing waves beyond the frictional boundary layer.

Therefore, this apparent bulge in the kinetic energy near the bot-

tom is not explained by the above theory. The possible existence

of small bottom slopes may be crucial here. The above theory as-

sumes a flat, horizontal bottom. Wunsch (1969) has shown that if

the group velocity vector for the reflected wave has an inclination

angle to the horizontal which is close to the bottom slope, an in-

tensification of the internal wave energy near the bottom can take

place. Since the group velocity vectors for waves near the iner-

tial frequency are only slightly inclined to the horizontal, the

near-bottom behavior of these waves .may depend strongly on small

bottom slopes.

'--- - -I~-~IC PIIPIPlllr~U P WI~-~ir* X -~ I *XI*Ui-I~~.*-III .~LII-_ PrD

Page 89: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

Chapter V Comparison of Vertical Wave Number Spectra with Theoret-

ical Models. Computation of Dropped, Time-lagged, Rotary Coherence

(DLRC). The Vertical Energy Flux of Internal-inertial Waves.

a.) Vertical wave number spectra

Recently, several models have been derived which describe

the amount of energy contributed to a region of the ocean by in-

ternal waves of various length and time scales. Perhaps the pri-

mary model of this type is that of Garrett and Munk (1972, 1975).

These authors have constructed an internal wave energy density

spectrum, E(a, w), as a function of horizontal wave number, a, and

frequency, w. The energy spectrum is assumed to be isotropic;

that is, the magnitude of E is independent of the direction of a.

The general form of E(a, m) is (Garrett and Munk, 1972, eq. 6.3):

E(a,w) = C1-A(X)2(w). V-1

This is a dimensionless equation. Time has been made dimensionless

by a reference Brunt-Vdislla frequency,

N0 = 3 cph, V-2

and distance by a reference cyclical wave number,

M = 0.122 cpkm. V-3

C is a constant. p is called the "a-bandwidth of the spectrum" as

___~__Ufl__ X _ _____L____ _~~_li~-~^-

Page 90: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

a function of frequency. The parameter X is given by:

A = a/p . V-4

The form chosen for 2(w) was:

= -p+2s 2 -s V-5

where wi is the (non-dimensional) local inertial frequency. The

constants p and s were chosen to be 2 and 1/2, respectively. Then

1/ 2 V-6

In the 1972 version of the model, A(X) was choien to have the form

of a "top hat":

A(X) = 1, for 0 < X < 1

and A(X) = 0, for X > 1. V-7

Finally, p(w) was chosen to be:

() = j.(W- .2)I/2 V-8* 1

where j, is an integer. In the 1972 model, j, was called the

"equivalent number of modes at the inertial frequency" by the

authors. (It should be emphasized that their model does not really

depend on the existence of modes in the vertical. The parameter j,

dan be thought of as characterizing the maximum vertical wave number

found to contain significant energy in the internal wave energy

Page 91: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

density spectrum). We would now like to see how this theoretical

model compares to vertical wave number spectra observed by the EMVP.

In order to proceed further we must introduce the dimension-

al forms of the above equations. In what follows, a hat (^) above

a quantity will indicate that it is dimensional. A prime (') will

indicate that vertical wave numbers in the stretched coordinate

are being used.

The dispersion relation for internal waves (under a WKB

approximation) gives:

a = i(z) (2 2 for 2 <)2

A(z) V-9

where ii is the local vertical wave number. Also,

= j*r (W2 i2) .)- = j*r NV-101 N V-O

Then (z) NX = -lJ

= - =

1M (j*) MO (z) V-11

However, since in the stretched vertical coordinate

m' = i(z)(No/N(z)), V-12

we have

(j,) Mo V-13

Page 92: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

Then the dimensional form of n(w) becomes

NS(G) = -,

8(A2 _ f2)1/2 V-14

and

N3E(W,m') = C A(X) o

(j.") e(^ - f2) . V-15

The total energy in horizontal velocity is given by

Etotal = E(a,w) WF d~ad V-16

whered2 + W.2

N wo (Garrett and Munk, 1972). V-17

However, since velocity in the profiles has been normalized by

N(z)/N o , we must use

U21

(A(z)/N o ) V-18

instead of U2 in equation V-16. Converting a and w to their

dimensional counterparts, where

A A

a = a/Mo and w = W/N , V-19

we have

E U2' E(a,w) da d' V-20total M0N 0

Page 93: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

Instead of integrating over ^, we wish to carry out the integration

over ml. Since

d= 2 - 2)/2 d V-21R(z)

and

d' = (No/i(z)) d, V-22

the integral becomes

Etotal 1 U2' E(i',) (2 1/2 dm ' d

NoMoo M NoA f

Co (,) (-2 +' in' d0 m' 0 W No

S o A( ) dw dm' . V-23(j*7)Mo G 3(-2 - 2)1/2 j Mo

Etotal must be multiplied by PNoMo - 3 to obtain the dimensional

energy spectrum, where P is sea water density (assumed equal to

1 gm/cm3 ). Removing the integral over i' leads to the following

dimensional form of the energy density spectrum in stretched verti-

cal wave number, Mi':

A A CN o ^E(m') = A + f2 dE-,($') = (-M-- ^/)*mIM o j'.M ^ ( _ f2)/2

No2 3Eo _i'

S- A ( )M0 2j*T j. Mo V-24

( C. Garrett, 1974, personal communication).

Page 94: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

E is a dimensionless constant which is related to C by

E = C l(w) dw V-25

To obtain the actual energy levels at a given depth, zo Em')

is multiplied by (N(z)/N ). At that level

m' = (zo) (No/N(zo)). V-26

The quantity EA,(m'), as obtained by observations, has been

given already in Figures 12 and 15. Clearly, the form of the

stretched vertical wave number spectrum depends on A(x). In the

1972 model, A(X) was chosen to be a top hat. Comparing equations

V-7 and V-13, we see that A(X) = 1 for m' < j,* THo and A(A) = 0 for

m' > j,Mo. (Again we see that in the top hat model j* controls

the maximum vertical wave number that contributes energy to the

spectrum). This means that the theoretical Em, is independent ofA

m' for X < 1, and is zero for X > 1. It is clear from Figures 12

or 15 that E~, is not independent of ii', but in fact decreases withm

increasing m'.

In a revised form of their model (Garrett and Munk, 1975),

these investigators chose a new form for the function A(X):

A(X) = (t - 1) (1 + X) -t V-27

where t is chosen to fit observation. Garrett and Munk picked

t = 2.5.

1__111___~_______4_1Y_~

V-28

Page 95: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

(The choice of t = 2.5 appears to agree reasonably well with verti-

cal displacement spectra obtained by Hayes, et. al.(1975)). For

No = 3 cph, Mo = 0.122 cpkm, Eo = 27 x 10 5 and j, = 6 (in the

1972 model, j. was 20), the vertical stretched wave number spectrum

becomes

2,(i') = 19.1 A(X) (cm/sec)2

m cpkm V-29

and

X = 0.435 i' . V-30

We have repeated the stretched vertical wave number spectrum

of Figure 12 in Figure 23. Figure 23 also shows the theoretical

stretched vertical wave number spectrum, E^.(m'), using the form of

A(X) given by equation V-27, for three choices of j.: j, = 6,

j. = 15 and j, = 20. (The cyclical frequency has been converted

from cycles per kilometer to cycles per sdb). The value of the

constant for the spectral energy level, E(j.), (corresponding to

Eo in the above derivation) has been adjusted to make the three

theoretical curves have the same energy levels at large stretched

vertical wave numbers. If E(j*) is the actual value of the con-

stant used for a given j., and Eo = 21 x 10 5 , we have the follow-

ing values for E(j,): E(j, = 6) = 2 x Eo; E(j, = 15) = 0.6 x E0 ;

E(J. = 20) = 0.4 x Eo. The theoretical curve for a j, of 15

appears to best represent the observed stretched vertical wave num-

ber spectrum, both in the general shape of the spectrum and in the

total energy level.

Page 96: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

WAVE LENG TH(SDB)

100 20

0.01 0.05

VER TICA L WA VE NUMBER(C/SDB)

Figure 23. Comparison of theoretical and observed total energy spectrain stretched vertical wave number (see text).\

Kj

(i)

Oz

(3S

(r~

(31

102

10

--lll~li~_~~ I__ ^1 _I*__I.~ I ~ _ I__LIII_1IIYI

Page 97: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

A value of 15 for j, is larger than the value of 6 assumed

in the 1975 model. If we return for the moment to a consideration

of Figures 10 and 11, the reason why j, = 6 is apparently too small

becomes clearer. Visually, Figures 10 and 11 show that waves con-

taining relatively high energy have wave lengths as short as

100 sdb. The total depth in the stretched vertical coordinate is

1860 sdb. If we now assume that j, is in some sense representative

of the greatest number of stretched vertical wave lengths that

"fit" into 1860 sdb (and contain significant energy), we find that,

for 100-130 sdb vertical wave lengths, j, should be at least 15.

It is also interesting to note that the value of E required

to make the theoretical (j, = 15) and observed stretched vertical

wave number spectra contain (roughly) the same total energy is

about 60% of the constant, Eo, chosen by Garrett and Munk. The

fact that high-frequency internal waves (those with frequencies

near N) are probably not energetic enough to be sensed by the EMVP

may partially explain this. However, it is also likely that the

total energy of the internal wave field at the MODE central moor-

ing is less than the total energy at other locations (at Site D

(390 20' N, 700 00' W), for example). Since observations at Site D

were used (at least in part) to determine Eo, the fact that

E(j, = 15) is less than E0 may also reflect an actual decrease in

internal wave energy at the central mooring, relative to other

locations in the ocean where measurements have been made.

~I.I_-La~L~XXI~II^WL.1 I~Y~

Page 98: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

b.) Dropped, time-lagged, rotary coherence of profiles

The following calculations were carried out in order to

compare profiler horizontal velocity data with some theoretical re-

sults based on the Garrett and Munk model presented in section a.

(C. Garrett, 1974, personal communication). Since we have a time

series of profiles, it is possible to calculate the velocity coher-

ence between.two profiles separated by a time interval, T, for any

stretched vertical wave number. This is called the dropped, lagged,

coherence (or DLC) by Garrett and Munk. However, we know that much

of the variability in the profiles is associated with waves having

periods near the local inertial period. For this reason, instead

of the DLC we have calculated the coherence between two profiles

when the horizontal velocities of the second profile have been

rotated through an angle y:

y = OT,

W = 2rrf, V-31

where f is the inertial frequency and T is the time interval be-

tween the two profiles. This is called the dropped, lagged, rotary

coherence (or DLRC). If the first velocity profile is

W,(i ) = Ul(u ) + iV1(z) V-32

and the second profile is given by

W2 (*) = U2 (*) + iV2( ) V-33

then the coherence is calculated between W1 and

WS = W2e = U3(**) + iV3(i*) V-34

_L---l-~- IIY~ -II-IYII~YX~~IIII IY --̂ Xi-I III-I~L~L~-- -.l11 ~rrrm-~-.

Page 99: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

We then obtain the following four quantities:

UU3 PUlU3PUlU3[PUlUl 1 /2[PU3U3]I/ 2 '

PVlV3 = PVIV3

[PVlVI] 1 /2[PV3V3] 1/2 '

QUIU3 = QU1U3

[PUlU] 11/ 2 [PU3U3] 1/2

QV1V3 = QVlV3

[PVIV ] 1/ 2 [PV3V3] 1 / 2 , V-35

where PUIU3 is the normalized cospectrum between velocity compo-

nents U1 and U3, QUlU3 is the normalized quadrature spectrum between

the same components, and PVIV3 and QVlV3 are the same quantities

for the velocity components VI and V3. PU1UI, PU3U3, PVlVl and

PV3V3 are the autospectra for U1, U3, V1 and V3, respectively.

All of the above quantities are functions of stretched vertical

wave number and the time lag between profiles.

We now describe the actual method of computation. Time lags,

T, were calculated for all pairs of drops in the time series. These

pairs were then sorted in such a manner that all pairs with lags

within + 1/2 hour of a given whole hour were assigned a lag equal

to that whole hour. For example, all pairs with time lags of from

2 1/2 to 3 1/2 hours were given a lag equal to 3 hours for the cal-

culation. Calculations were done only for those lags which had

four or more pairs of drops. The results of these calculations are

Page 100: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

100

shown in Figures 24, 25, 26 and 27 for lags up to almost two days.

For any lag at which the calculation was made, there are three

points shown for each of the four normalized spectral quantities

given above. Each of these points represents an average over all

pairs available at that time lag. In addition, estimates have

been averaged over several stretched vertical wave number bands.

The number n represents the number of the stretched vertical wave

number estimate; that is, the vertical wave length is given by

1800/n sdb, where 1800 sdb is the total length of the series.

Thus, the three points (at a given lag and for a given quantity)

are averages over stretched vertical wave number estimates n = 1-5,

n = 6-10, and n -, 11-15.

It is somewhat difficult to put error bars on PUlU3, PVIV3,

QUIU3 and QVIV3 since each point may be the result of averaging

over a different number of pairs of drops. In an attemipt to esti-

mate the error we have calculated PUlU3, VlV3, QUlU3 and QVlV3

separately for each pair of profiles in a case when four pairs of

profiles were used to give the estimates shown in Figures 24-27.

(We recall that the smallest number of pairs used was four.) The

resulting values of PUIU3, PVlV3, QUIU3 and QVlV3 have an average

standard deviation about their mean values of about + 0.2. Pre-

sumably, points which are obtained by averaging over more than four

pairs of drops are even more stable; that is, the scatter of points

about their mean values is less than + 0.2 standard deviation.

Page 101: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

10

T (HOURS)20 30 40 50

Figure 24. Dropped, lagged, rotary cospectrum (normalized) between velocity componentsU, and U3 (see text for a further description of Figures 24 to 27).

Page 102: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

T (HOURS)20 30

Figure 25. Dropped, lagged, rotary cospectrum (normalized) between velocitycomponents V, and V3.

10 40 50

t)

Page 103: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

T (HOURS)10

t20 30

* n=I-5o n = 6-IOx n = 11-15

0

xoxg x

Xoo00;6

x x0 o

Figure 26. Dropped, lagged, rotary quadraturevelocity components U, and U3.

spectrum (normalized) between

40 50

<0O

I

L

Page 104: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

T (HOURS)20 30

Figure 27. Dropped, lagged, rotary quadrature spectrum (normalized) betweenvelocity components V1 and V .

10 40 50

Page 105: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

105

Figures 24 and 25 also show the curve of PUIU3 or PVIV3 ob-

tained theoretically (C. Garrett, 1974, personal communication) from

the Garrett-Munk model under the assumption of an isotropic spec-

trum. The curve for PUIU3 or PVIV3 does not depend on wave propa-

gation in the vertical. In contrast, UlU3 and QVIV3 do depend on

vertical propagation. If only vertical modes exist, QU1U3 and QV1V3

are theoretically always equal to zero. On the other hand, if all

waves are propagating energy downward, the theoretical curve for

QUlU3 or QVIV3 starts out at zero for zero lag and slowly increases

to about + 0.25 for a lag of 50 hours. If we accept the above

"confidence limit" of + 0.2, it is difficult to say whether the ob-

served values of QUlU3 and QVIV3 are significantly different from

zero. Therefore, the following paragraphs will be limited to a dis-

cussion of the observed behavior of PUlU3 and PVIV3. These quan-

tities appear to contain the most interesting information on the

behavior of the DLRC with varying time lag and stretched vertical

wave number.

Using the above error estimate, it is possible to say sev-

eral things about PUlU3 and PVlV3. First, the behavior of PUIU3

or PVIV3 for estimates containing the largest vertical wave

lengths (n = -1-5) is significantly different from the behavior of

these quantities for shorter vertical wave lengths (n = 6-10 and

11-15). The estimates for n = 1-5 decrease much more rapidly with

increasing lag than the estimates for higher n. In fact, the ob-

served PUlU3 and PVIV3 curves for n = 1-5 have an oscillatory

Page 106: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

106

character with minima (greatest negative values) for lags of about

12 hours, and again for lags of about 30-40 hours. This is a

strong indication that the longest vertical wave lengths have sig-

nificant energy contributions from waves with periods near 12 hours.

The most obvious candidate for this energy contribution is, of

course, the barotlinic semidiurnal tide. In contrast, the esti-

mates for higher n show little evidence of energy contributions at

approximately 12 hour periods. This implies that most of the

semidiurnal tidal energy is in the longest vertical wave lengths.

Other work (Hendry, 1974) has shown that the baroclinic, semi-

diurnal tide has most of its energy in the lowest three vertical

modes. The above result appears to confirm this.

Second, PU1U3 and PVlV3 for higher n (n = 6-10, 11-15)

follow the theoretical curve more closely. Given the above "error"

estimate, PU1U3 and PVIV3 curves for n = 6-10 and n = 11-15 cannot,

in general, be distinguished. The theoretical carve does not de-

pend on stretched vertical wave number, but the error is probably

large enough so that dependence on vertical wave number cannot be

tested adequately. It is interesting to note that PU1U3 and PVIV3

for n = 6-10, 11-15 are almost always greater than the theoretical

curve. -A possible explanation for this is that while the theoret-

ical curve takes into account all internal wave frequencies (from

inertial to Brunt-Vlisglg), the sensitivity of the profiles is such

that higher frequency internal waves are "filtered out." These

higher frequencies tend to make 1U1U3 or PVIV3 decrease more

Page 107: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

107

rapidly with increasing lag than would be true if only near-inertial

waves were being observed. The results of this section may be sum-

marized as follows: 1.) The behavior of PU1U3 and PVIV3 for wave

lengths outside of the region in which semidiurnal tidal energy is

significant does not appear to refute the theoretical calculations,

taking into account the possible sources of error described above.

2.) The result of other calculations, showing that most of the

semidiurnal tidal energy is confined to the largest vertical wave

lengths, is also obtained from the present calculations.

c.) The vertical energy flux of near-inertial waves

We have pointed out that, because of the relatively short

time interval over which the time series was obtained, it is dif-

ficult to obtain good frequency resolution for the observed waves.

However, there are several independent characteristics of the data

which lead us to believe that the waves which we have been calling

"inertial" actually have periods somewhat shorter than the inertial

period.

First, Figures 6 and 7 indicate that the phases of the domi-

nant waves are moving upward in time. If these waves were of ex-

actly inertial frequency, internal wave theory indicates that the

phase should remain at the same depth, independently of time. In

order to propagate upward, these waves must have frequencies greater

than f.

Page 108: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

108

Second, consideration of Figure 18 shows that many of the

more energetic velocity features which can be identified between

drops 235U and 236U (these drops being separated by 4.8 kilometers)

do not occur at the same depth in both profiles, but rather are

offset slightly. This depth offset is too large to be accounted

for by errors in the EMVP pressure calibration. The implication is

that the phase planes of these waves are tilted slightly away from

horizontal. Figure 21 then indicates that the wave number vector,

, for these waves must be tilted at a small angle to the vertical,

and equation III-10 shows that the wave frequency, 6, has to be

slightly greater than f.

The third indication that the wave frequencies are somewhat

greater than f comes from attempts to least-squares fit sinusoidal

waves to the data. At any pressure, z, it is possible to fit sinu-

soids of the form:

ufit (Zn,ti) = A(zn) sin((zn) + ti) ,

vfi t ("n,ti) = B(in) sin( (zn) + sti) V-36

over a group of profiles. t i is the time of a given profile. u and

v are east and north velocity components. Then the quantities

U() = . (u (n ,ti) - Uobs (Z ,ti))2 V-37

and

V(n) =i f i t( i) - vob s n'ti))2 V-38

Page 109: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

109

give an estimate of the residual error in the fit at depth zn.

These fits can then be calculated over a range of depths, Zn. The

quantity

D = E U() + V(zn) V-39zn

represents a measure of the total residual error of the fit over

the time and depth interval studied. Attempts to do such fits in

the main thermocline, where the strongest signal is found, indi-

cate that the total residual, D, is minimized if the period, T,

used in the above formulas (' = 2r/T) is slightly less than the

inertial period. In fact, the best fit occurs if we choose u so

that T = 22-23 hours. The inertial period is 25.56 hours at the

latitude of the profiler time series.

We have argued in previous chapters that the near-inertial

waves are propagating their energy downward from the surface. Since

waves with frequencies close to f have group velocity vectors that

are only slightly inclined to the horizontal, it is at first glance

surprising that these waves appear to make significant contributions

to the energy even in the deep water. It could be argued that those

waves which have the maximum (downward) vertical energy flux will

be able to penetrate from the surface to the deep ocean most effi-

ciently. This vertical energy flux is given by

F = Cg E V-40

where Cgz is the vertical component of the group velocity vector,

Page 110: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

110

and E is the total (kinetic and potential) wave energy. It is of

interest to obtain the local tgz, as a function of z, and use it,

along with the theoretical energy spectrum of the previous section,

to calculate F as a function of frequency and vertical wave number.

Because the derivation of the Garrett-Munk spectrum assumes

that a WKB-type description of the wave behavior is adequate, we

will assume in what follows that this is again true. The equations

to be solved are the same as equations B-6, except that now N is a

function of depth. The equation for the vertical velocity, w, is:

d2dw + 2 2 ,2) w = 0 V-41dz2

with

k;2 +22 =

S - 2 . V-42

To obtain the WKB solution, we let

8 (z)dz 0 A n(z)W= e E n V-43

n=O n V-4

The solutions for n = 0 and n = 1 give, respectively, X and Ao -

2t[a 21/2X = + i [N (z) - w2] V-44

and

= 10 A 2 1/4 V-45

[N -w 2Then

21/2+ i k - 12 2 2 ) dz

1 e- zo2.sE[2(z) - 2]1/4 V-46

Page 111: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

111

where the region of integration is over an interval of z for which

N2(z) > ^2. From continuity ,we get u and v:

u = ( ok ifl) 2 1/4 + +B X(z)dzA f(N2(Z) _ 2} 1z/

v = (t iR (2 _ f 2) /2 (k2 + ~2) V-47

The buoyancy is

1 xdz

b= e

i( 2 2) 1/4 V-48

The total energy, E, is given by

* = 1/2 ( < j j2 + I^12 + I -2 )^ ^ ... 1^12 + A12 + 1-1 2 +

2 1/21/4 (N2 - 2) 2 +2 +N2

=21/4 _ ^2 2 2 _ V-49

where the angle brackets denote an average over time. The ratio

of vertical kinetic energy plus potential energy to horizontal

kinetic energy is:

vert. K.E. + P.E. (N2 + W2) (2 f2 )

horiz. K.E. (i - W) (w2 2) V-50

For frequencies close to i, almost all of the energy appears as

horizontal kinetic energy. The vertical component of the group

velocity vector at depth z isgz(8,,z) = = - 2 ^2) (^2 _ 2)

a mu(z) ( - 2)V-51

Page 112: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

112

The theoretical spectrum of Garrett and Munk has already been given

(equation V-24). This spectrum contains only horizontal kinetic

energy. However, equation V-50 shows that for frequencies close to

, most of the wave energy is horizontal kinetic energy. From

equation V-24, the energy density spectrum is

N 2

m ( m'',0 ) = BC -- A (--+) +f2Mo3 j,Mo ~ 1M 3(2 _ 2)1/2 V-52

The vertical energy flux for waves in a small interval of frequen-

cy-stretched vertical wave number space Am'AW is

Fz = Cg N (m',) Am' , V-53Fz Cgz NO

where m' and ^ are taken to be in the middle of the element Am'Aw.

The vertical group velocity can be expressed in terms of w and m'

by using equation V-12

^ ^ N ( - f2) (^ - 2)Ceg ( ,m',z) = -

z(z) , (R2 -2) V-54

Then2 1/2NC 2 (W2 + f2)-(N2 w ~2) (W2 2)1/2 #% '

Fz = - C-A M .Mo Mo ^ 2 - 2)

V-55

We note that for w close to f, Fz is almost independent of z.

From the WKB approximation, it can be shown that

U aS= z (Cgz =0. V-56a

Page 113: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

113

However, since we are only considering horizontal kinetic energy,

aFz/az is not equal to zero but approaches zero as w approaches f.

We are particularly interested in examining Fz as a function

of w, for fixed m'. The frequency dependence of Fz is given by

g() = (2 + f2) (^2 - ^2)1/ 2(2 - 02)

w (N _ 2

(o2 + f2) ( 2 2)1/2(1 - 2

a' V-57

since N >> f. If W and N are made non-dimensional by

f f V-58

then

(W + 1) (W 2 1)1/2(1 -0)2

2

W V-59

f() is plotted in Figure 28, with N = 5f. It can be seen that

f(') increases rapidly with increasing ' until ' = 1.2, and then

decreases. This implies that, for fixed m', Fz reaches a maximum

for ^ = 1.2f. For an inertial period of 25.56 hours, the vertical

flux of horizontal kinetic energy reaches a maximum (at a given M')

for waves with periods of around 21 hours. This period appears to

be somewhat shorter than that which might have been expected. (We

recall that the least-square fit indicated that the dominant waves

in the profiles seemed to have periods of 22-24 hours in the main

Page 114: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

I .0

0.5

0.0-2 3 4 5

Figure 28. Frequency dependence of the vertical flux of horizontal kinetic energy, usinga theoretical form for the energy spectrum of the internal wave field.

Page 115: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

115

thermocline.) However, it is significant that the downward energy

flux appears to have its maximum value close to the local inertial

frequency in this model.

A certain amount of inconsistency is inherent in using a

spectrum such as the Garrett-Munk model to calculate energy flux.

The first point of inconsistency, of course, is that the true spec-

trum will not. have an infinity at the inertial frequency. This

problem could be removed by assuming, for example, that there should

really be an energy cut-off near the inertial frequency; that is,

instead of the infinity at the inertial frequency, the spectrum

could be altered locally (at and near f) so that a simple peak oc-

curs at the inertial frequency. This alteration would not change

the flux curve of Figure 28 appreciably. In fact, it would force

the observed flux maximum to be concentrated even more about per-

iods of 21-22 hours. The second problem is more subtle, and in-

volves the question of why one would expect the spectral peak to

be exactly at the inertial frequency. In a case (such as the

Garrett-Munk model) where vertical modes have been assumed to have

been established a priori, it seems that a peak at the inertial

frequency would not be inconsistent, since the waves are not re-

quired to be supported by energy sources located, for example, at

the surface or bottom boundaries. However, once we require that

the waves be supported by a flux of energy into the ocean from the

surface (or bottom) boundary it is not consistent to expect that

the peak in the observed energy spectrum at some depth will be

Page 116: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

116

found at that frequency (f) where the vertical energy flux becomes

zero. Both of the above points argue that the "inertial peak"

should actually appear at frequencies somewhat greater than f.

Figure 29 shows a frequency spectrum obtained from averaging

spectra of current meter records at 1500 meters at 12 different

locations in the MODE region (M. Briscoe, 1975, personal communi-

cation). The-average inertial frequency of the 12 locations is

also shown. (The maximum and minimum inertial frequencies of the

12 locations are about + 0.002 cph about the average inertial fre-

quency.) It is clear from Figure 29 that the inertial "peak" is

displaced toward frequencies greater than f. The center of this

wide peak is in the period range 22-23 hours, in agreement with the

above discussion.

It is possible to use the above expression for the vertical

group speed (equation V-54) along with the observed stretched verti-

cal wave number spectrum (Figure 16) to calculate a rough estimate

of the vertical energy flux (downward and net), integrated over all

vertical wave numbers. In the absence of accurate frequency infor-

mation, we have assumed that the clockwise and counterclockwise

spectra are dominated by waves with periods of the calculated max-

imum flux (or of the center of the near-inertial "peak" in Figure

29); that is, periods of 22-23 hours. Using this period range in

the expression for the vertical group speed, and assuming that

roughly half the energy in the clockwise and counterclockwise energy

estimate for the smallest two or three stretched vertical wave

Page 117: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

117

.. 00' IO7'cr.

FREQUENCY ICPN)

Figure 29. Frequency spectrum, obtained by averaging 12 spectra computedfrom current meter records at 1500 m in the MODE region. Energydensity units are (cm/sec)2/cph. f is the inertial frequency.

Page 118: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

118

number estimates is contributed by the semidiurnal tide (see Chap-

ter V-b), we obtain estimates of the total downward wave energy

flux of 0.6-1.0 erg/cm 2/sec, and of the net downward wave energy

flux of 0.2 - 0.3 erg/cm 2/sec. The estimate of the total down-

ward flux is consistent with other estimates of vertical energy

flux by internal waves (Bell, 1975). We have suggested in pre-

vious chapters that a bottom boundary layer could be a significant

energy sink for near-inertial waves. An order-of-magnitude cal-

culation indicates that a net downward energy flux of 0.2 - 0.3

erg/cm2/sec could be matched by energy loss in a boundary layer.

A measure of this possible energy loss is

L = 4K p((u)2 + (A )2) Az erg/cm2 /sec V-60v Az Az(Yih, 1969)

where Kv is the coefficient of vertical eddy viscosity, p is den-

sity, u and v are horizontal velocity components, and Az is a scale

depth of the boundary layer. Choosing values for Au2 + Av2 and

Az from Figure 8 (Au2 + Av2 - 4 cm2/sec 2, Az ~ 5 x 10 cm), we get

L - (0.3 x 10-2) Kv erg/cm2 /sec. For Ky = 100 cm2/sec, L -~ 0.3

erg/cm 2/sec. Bell (1975) quotes a value for Kv of - 10 cm2 /sec in

the interior but notes that Kv could be at least an order of mag-

nitude larger in a turbulent bottom boundary layer. Thus, the

value of Kv needed to make L comparable to the estimated net down-

ward energy flux is not inconsistent with other estimates of Kv.

Page 119: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

119

Chapter VI Some Observations over Rough Topography

In previous chapters we have presented an analysis of data

obtained over smooth topography (at the MODE central mooring).

Some of the calculations presented, such as reflection coefficients,

depend on the assumption that the bottom is not acting as a wave

source. Over the smooth topography this does not appear to be an

unreasonable assumption. However, it is well known from theoretical

studies that a rough bottom may interact with either a mean flow,

producing lee waves, or with an oscillatory flow such as a tide.

Bell (1975) has discussed the possible generation of internal waves

in the open ocean by rough topography. In an earlier work,

Hendershott (1964) analyzed a case where inertial waves could be

generated by the oscillatory flow of a diurnal-tide component over

a rough bottom. His study indicated that strong inertial wave

currents might be generated at latitudes slightly to the south of

the critical latitude (the latitude at which the tidal and inertial

periods are equal). The critical latitude for the 24-hour diurnal

tide would be 300 N, for example. His theory actually predicts

that there should be a somewhat complicated band structure of the

inertial energy south of the critical latitude, with a primary peak

in the inertial energy just south of the critical latitude, and

secondary maxima farther south.

Page 120: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

120

It is important to realize that there are numerous tidal

lines that are called diurnal.' Although the critical latitude for

the 24-hour tide is 30* N, the critical latitude for the 24-hour

50-minute tide moves down to about 28.90 N. This means that the

actual energy levels of the inertial waves (generated by topography)

might not depend as strongly on latitude as is predicted by the

theory. In any case, the fact that the major diurnal tidal compo-

nents at 280 N have frequencies within the internal wave frequency

range, but near the inertial frequency, indicates that it would be

worthwhile to look for near-inertial energy propagating away from

the bottom in the rough topography region.

During the MODE experiment, a set of prcfiles was obtained

around a ridge in the rough topography region east of the central

mooring. These drops are indicated by a dagger (t) in Appendix A.

The center of the ridge is located at approximately 280 04' N,

68* 31' W. The main part of the ridge is oriented along the axis

NNE-SSW and rises about 700-800 meters above the adjacent, smoother

part of the ocean floor. Because this set was obtained in order to

make a survey of the velocity structure around the ridge it is not

a time series at one location. In fact, the set is made up of a

series of pairs of profiles, except in a few cases where only single

profiles were made. Each pair was obtained at one location around

the ridge, and the second profile was usually taken at about one

half of an inertial period after the first profile. Because a time

series of profiles is not available over the rough topography, it

Page 121: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

121

is difficult to accurately remove the low-frequency shear profile.

For this reason, the results of calculating clockwise and counter-

clockwise vertical wave number spectra are more difficult to inter-

pret than are the corresponding results over the smooth topography.

However, some significant features in these rough topography pro-

files can be seen visually; that is, without recourse to spectrum

calculations.

Figures 30 and 31, for example, show hodograph plots (plots

of the horizontal velocity vector with depth) for profiles 190D and

192D between 3000 dbar and 4500 dbar. These two profiles were made

at a time lag of one half of an inertial period. The original pro-

files have been smoothed by applying a running 150 dbar triangular

filter to each velocity component. This makes it easier to vis-

ually follow the behavior of the longer wave length (higher energy)

waves in the profiles. Smoothed points are plotted every 10 dbar,

and depth is marked every 100 dbar. The most striking feature of

these hodographs is that they show wave-like structures for which

the horizontal velocity vector is clearly rotating counterclockwise

with increasing depth. The structure is clearly time-dependent, as

can be seen by comparing the hodographs for profiles 190D and 192D.

Comparison of the hodographs appears to indicate that these waves

have periods near the inertial period. We have already pointed out

in Chapter III that such counterclockwise polarization indicates

that these waves are propagating energy upward, away from the ridge.

Page 122: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

NORTH (CAM/SEC)0ro)

CO'(V

0NO

|V

t)|

0.

-

L)6 I

1 L.

38

37

34

40

32

3630

4543

3.0-

2.5-

2.0-

1.5-

1.0-

0.5-

0.0-

-0.5

Figure 30. Smoothed hodograph of profile 190D between 3000 dbarand 4500 dbar. Depths are in hundreds of decibars.

4.5

122

4.0 -

3.5-

lLu

K)K;(I)

p 1-

O

d.

Page 123: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

EAS T (CM/SEC)I

0U' 06 001 I0

-t r IrI. _

34

33

5.0

4.5-

4.0-

3.5-

3.0-

2.5-

2.0.

30

31

1 I I I I I I I I

Figure 31. Smoothed hodograph of profile 192D between 3000 dbarand 4500 dbar. Depths are in hundreds of decibars.

5.5

123

32

K)

Q::(I)

1.,5-

i.U

Page 124: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

124

Although spectral estimates obtained from only two profiles

are probably not very stable, it should be mentioned that vertical

wave number spectra (without stretching the vertical coordinate)

obtained in the deep water (2000 dbar and deeper) for drops 190D,

1910, 192D and 193D.show a consistent dominance of counterclockwise

energy over clockwise energy. These four drops were made at the

southwest end- of the ridge.

In contrast, deeper water (greater than 2000 dbar) spectra

and hodographs for a drop pair, 195D and 197D, that was made north-

east of the ridge, do not show any strong tendency for counter-

clockwise rotation with depth. If anything, these drops show a

tendency for the velocity vector to rotate clockwise with depth in

the same manner as was observed over the smooth topography. Swallow

floats located near the ridge when these profiles were made indi-

cated that the mean flow (at 4000 dbar) was to the southwest.

Since the pair 195D and 197D was located to the northeast of the

ridge while 190D, 191D, 192D and 193D were to the southwest, we

might speculate that if the ridgewas generating near inertial waves,

these waves would be advected to the southwest and therefore would

be seen primarily by profiles to the southwest of the ridge. Al-

though it is difficult to draw general conclusions from a small set

of profiles, the above observations in the deep water around the

ridge can be summarized as follows: profiles made to the southwest

of the ridge show clear evidence of counterclockwise polarization

with depth; profiles made near the ridge crest (such as 189D, for

Page 125: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

125

example) do not appear to show a dominance of either polarization;

two profiles (195D and 197D) made over a smoother region to the

northeast of the ridge appear to have reverted to the clockwise

polarization with depth characteristic of the smooth topography

profiles, at least for the longer vertical wave lengths, which are

most energetic.

Bell (1975) has calculated a value for the possible upward

energy flux of internal waves generated by rough topography of

-~ erg/cm 2/sec. This value is comparable to the total downward

flux in the smooth topography region, as estimated in Chapter V.

If the average vertical energy flux values for bottom topography

and sea surface generation are comparable, it might be expected

that energy input at the bottom could sometimes exceed energy input

at the surface (especially since both input processes are probably

intermittent in time). This would result in waves polarized

counterclockwise with increasing depth, at least in the deep water

over the rough topography. (We have also examined hodographs of

the rough topography profiles in the upper half of the water

column. In contrast to the plots in the lower half, these hodo-

graphs do not show a clear dominance of one polarization over the

other, indicating that in the upper half of the water column a

complicated addition of the top and bottom energy sources may be

taking place.)

Page 126: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

126

Chapter VII Conclusions, a Further Discussion of Some of the Data

and Recommendations for Future Experiments

The previous chapters have been devoted to an analysis of

internal wave data obtained by vertical profiling. It is now

worthwhile to summarize the significant results of this analysis

and place these results in the general framework of internal wave

theory.

One of the more fruitful results of this analysis has been

the ability to relate the observed characteristics of wave polar-

ization and vertical phase propagation to the locations of energy

sources for these waves; that is, to the possible existence of

internal wave energy sources at the top and bottom boundaries of

the ocean. As Muller and Olbers (1975) have pointed out, a two-

sided, symmetric, vertical wave number spectrum (equal energy in

positive and negative vertical wave numbers) is not consistent with

the existence of net boundary sources at the surface and bottom.

A major cause of difficulty with internal wave observations until

now has been the inability to observe asymmetry of internal wave

vertical wave number spectra (or, in fact, vertical wave number

spectra) of horizontal velocity with the techniques that have been

employed. Time series of horizontal velocity obtained from current

meters at fixed locations in space cannot distinguish between an

internal wave having an upward or a downward phase propagation,

Page 127: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

127

since the horizontal velocity vector rotates in the same sense for

either wave (clockwise in the northern hemisphere). The tech-

nique, employed in Chapter II, of decomposing vertical profiles

into clockwise and counterclockwise energy contributions is a

useful method for observing asymmetry in the internal wave vertical

wave number spectrum and for inferring the direction of net verti-

cal energy flux from this asymmetry. A similar instance where

vertical wave polarization has been used to infer the direction of

vertical energy propagation in the atmosphere has already been

pointed out (Hines, 1966). A case where vertical polarization of

tidal currents was apparently observed in an analysis of current

records obtained from current meters on moorings has been reported

by Titov and Fomin (1971). We have mentioned in Chapter III, how-

ever, that the ability to identify clockwise and counterclockwise

energy with waves having upward and downward wave number vector

components (that is, with the two-sided vertical wave number spec-

trum) requires that the observed energy be primarily contributed

by waves with frequencies near the local inertial frequency. This

is true because individual waves are nearly circularly polarized if

their frequencies are near the local inertial frequency. If a

single wave with a frequency much greater than inertial were pres-

ent, for example, this wave would provide energy contributions to

both the clockwise and counterclockwise vertical wave number spec-

tra, since the wave would be elliptically polarized in the vertical.

The clockwise energy would still be greater than the counterclockwise

Page 128: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

128

energy if the wave were propagating energy downward. But the

true, two-sided vertical wave number spectrum of this wave would

contain energy only on one side of the spectrum (at the positive

or negative vertical wave number of the wave, depending on which

way the phase of the wave is propagating in the vertical).

Fortunately, the data presented and analyzed in previous

chapters indicates that the two dominant sources of energy in the

profiles are near-inertial waves (at all vertical wave numbers),

and the semidiurnal tide (at the smallest vertical wave numbers).

Inertial and diurnal tidal contributions to the profiles cannot,

however, be resolved because of the short length of time over which

the profile time series was obtained. The ave-age frequency spec-

trum at 1500 meters (Figure 29) shows that a large part of the energy

observed at this depth in the internal wave band comes from fre-

quencies near the inertial-diurnal tidal frequency, and from the

semidiurnal tide. The observed "mirror-imaging" of profiles (Fig-

ure 3) and the calculations of dropped, lagged, rotary coherence

(Figures 24 to 27), support this.

The observation that much of the variability in the profiles

was.caused by the near-inertial waves (and the semidiurnal tide at

the largest vertical wave lengths) led to efforts to evaluate quan-

titatively the total and net downward energy flux over the smooth

topography. Making assumptions about the.amount of energy contri-

buted by the semidiurnal tide at the smallest vertical wave numbers,

we obtained values for the total vertical energy flux downward of

Page 129: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

129

~ .6 - 1.0 erg/cm2/sec, and for the net vertical energy flux down-

ward of ~.2 - .3 erg/cm 2/sec. Although both estimates are subject

to uncertainty because of the presence of the semidiurnal tide,

neither estimate seems unreasonable (see Chapter V-c). In partic-

ular, it appears that the net downward energy flux can be balanced

by a rough estimate of the rate of energy loss per cm2 in the bot-

tom boundary layer (if we assume, of course, that the observed de-

crease of mean perturbation horizontal kinetic energy near the

bottom is actually caused by the presence of a boundary layer).

In relation to the contour plots (Figures 6 and 7), a fur-

ther statement can be made which, although somewhat-obvious, is no

less important. The observed relations among wave period, direc-

tion of vertical phase propagation, and vertical wave polarization

in Figures 6 and 7 constitute at least a partial verification of

the dispersion relation for internal waves (those near 'the inertial

frequency). Other observations and analyses of data (from current

meter records, for example) have assumed that the observed oscil-

lations between the inertial and Brunt-Viisdl' frequencies obey

the dispersion relation for internal waves. In the absence of clear

evidence of the propagation of oscillations in this frequency

range, other investigators (Webster, 1969, for example) have ar-

gued that some other process (such as turbulence) could explain

the observed oscillations. Although turbulence is difficult to

define, one of its characteristics is that it does not follow a

well-defined dispersion relation. In this sense, then, the results

Page 130: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

130

of Figures 6 and 7 clearly argue for internal waves as being the

cause of the observed variability in the profiles, at least for

the near-inertial waves.

Another significant result brought out by this analysis is

the fact that both profiler data and other, independent observa-

tions (Figure 29) have shown that the energy concentration in what

one might call the "inertial peak" is, first, not concentrated in

a well-defined peak (as is the semidiurnal tide, for example) and

second, is not concentrated at exactly the inertial peak but is

rather centered at a frequency about 20% greater than the local

inertial frequency. Some previous observations have also shown that

the inertial "peak" is quite often displaced toward frequencies

somewhat greater than the local inertial frequency. Webster (1968),

for example, shows a frequency spectrum of current at 38* 28.8' N,

70* 00.5' W, in 3300 meter-water, with the observations made about

30 meters above the bottom. This spectrum has a clear concentra-

tion of energy near the local inertial frequency, but the center

of this peak occurs at a frequency about 5% higher than the local

inertial frequency. White (1972) considered temperature data ob-

tained in the upper several hundred meters at a site in the Pacific.

His temperature spectra always show an inertial "peak" at frequen-

cies up to 30% higher than the local inertial frequency.

Various attempts have been made to explain this shift theor-

etically. Munk and Phillips (1968), for instance, argued that

internal waves generated to the south of a given latitude of

Page 131: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

131

observation (in the norther hemisphere) would propagate northward,

and that as they propagated northward their frequencies would ap-

proach the local inertial frequency. A test of this idea by the

use of profile: data would require better information on the direc-

tion of horizontal phase propagation for the observed waves. We

will return to this question later, when the possible influence of

the mean vertical shear on the propagation of the waves seen in

the profiles is discussed.

White (1972) pointed out that the observed frequency shift

of the inertial waves could be explained as the result of a

Doppler shift due to the mean flow. Because he always observed a

shift toward higher frequencies, he was forced to conclude that

the near-inertial waves he observed were always propagating in the

direction of the mean flow. The data presented in previous chap-

ters, however, show that it is unlikely that Doppler shifting could

be the cause of the observed frequency shift. The average fre-

quency spectrum at 1500 meters (Figure 29) shows that there is an

appreciable (20%) shift of the "inertial peak" toward higher fre-

quencies. But it is known from a calculation of the average flow

observed by a current meter at 1500 meters near the location of the

five-day time series (Chapter II) that the mean flow at that depth,

when the time series was made, was small (< 3 cm/sec). It does not

seem that such a small mean flow could account for the observed

frequency shift. This does not necessarily negate White's conclu-

sion, since he was observing at a point where the mean flow was

Page 132: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

132

about five times that observed at 1500 meters in the present exper-

iment. It would seem fortuitous, however, that the near-inertial

waves would always propagate in the direction of the mean flow,

especially if, as the present data indicate, they are geing gen-

erated by atmospheric forcing at the ocean surface. In fact, there

is a suggestion in the data (discussed below) which indicates that

at least the more energetic features in the profiles have a compo-

nent of propagation to the north. Since the mean flow, at least to

about 1500 meters, was to the south during the present experiment,

this would imply that, if anything, the inertial peak would be

shifted to lower frequencies.

The results of the profiler observations now allow us to

propose another mechanism to explain why the observed "inertial

peak" frequency shift almost always appears to be toward higher

frequencies. This explanation is that the observed waves are

clearly propagating in the vertical. The contour plots (Figures

6 and 7) show that the dominant structures associated with the

near-inertial waves are moving upward in time, corresponding to

downward energy propagation. (Of course, these contour plots are

influenced by the mean flow. However, we have already pointed out

that around 1500 meters the mean flow is small, and yet the great-

est speed of upward phase propagation in these contour plots is

found in the 1500 dbar to 2500 dbar range.) In order to propagate

vertically, these waves must have frequencies greater than the

local inertial frequency. The calculations of Chapter V-c showed

Page 133: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

133

that, using an assumed form of the internal wave energy spectrum,

the maximum vertical flux of horizontal kinetic energy occurs at

frequencies about 20% higher than the inertial frequency. We

pointed out that, once the energy sources are placed at the surface

(or bottom) boundary, it is inconsistent to have the peak in the

energy spectrum appear exactly at the inertial frequency when waves

of this frequency cannot propagate into the interior. It is con-

sistent, however, to have an "inertial peak" that is shifted to

higher frequencies. It seems likely that the waves which, when

generated at the surface, have the greatest downward energy flux

would be most able to penetrate through the total water column.

(They would presumably be able to travel a greater distance in the

vertical before losing an appreciable fraction of their energy due

to friction, non-linear interactions, shear interactions, or some

other cause.)

We have argued that a significant part of the energy put in

by the atmosphere is lost in a bottom boundary layer, at least for

the near inertial waves. Calculations of the net downward energy

flux and the rate of energy loss in what we are interpreting in

Figure 8 to be a bottom boundary layer show that (in terms of a

rough calculation) the energy flux into the boundary layer could be

accounted for by the energy loss in the boundary layer. It is cer-

tainly true that other factors can act to remove energy from the

wave field. For example, it is clear that the observed waves tend

to increase in amplitude and decrease in vertical wave length in

Page 134: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

134

the thermocline, where the Brunt-Viisili frequency is large. For

waves shorter than some critical value, the shear caused by these

waves will be large enough so that the Richardson number would be

less than 1/4, and the flow caused by the wave might become unstable.

This would extract energy from the waves (possibly putting that

energy into turbulence). Such a process would be most active at

the high-vertical-wave number end of the vertical wave number spec-

trum. It is also reasonable to expect that frictional losses will

be important throughout the water column (not just in a bottom

boundary layer) for waves with short vertical wave lengths. As the

vertical wave length decreases, the frictional term; v 32 /dz2, in

the horizontal mrmentum equation (Chapter IV) for near-inertial

waves grows. In terms of the boundary layer discussed in Chapter

IV, we can say that (assuming the frequency is fixed and close to

inertial) as the vertical wave length decreases, the horizontal

wave length will also decrease. This will cause the parameter

R = vk2/f of Chapter IV to increase. When this occurs, frictional

losses are no longer confined to a thin bottom layer, but rather

become important in the interior of the water column as well. This

indicates that the argument that near-inertial waves are losing most

of their energy in a bottom boundary layer is probably only appli-

cable for near-inertial waves with small vertical wave numbers.

This process of energy loss due to friction in the interior of the

water column would also be expected to be most active at the high-

vertical-wave number end of the vertical wave number spectrum.

___L__XIL~___I I;__C- .-̂I-I~1IYII~-LII~-I~ -~i~lL

Page 135: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

135

Another cause of energy loss by these waves would be by

interaction of the waves with the mean flow. In processing the

profiler data we chose to normalize the profiles by the mean Brunt-

Vais5li profile. This type of normalization was used because it

made the calculations easier to compare with existing theories.

However, it is well known (Phillips, 1966) that a vertical shear

in the mean horizontal velocity profile can also cause changes in

the amplitude and wave length of internal waves as these waves

propagate through the shear profile. Perhaps the greatest diffi-

culty in evaluating the influence of the shear on the profiles is

lack of data on the direction of horizontal phase propagation of

the waves relative to the mean shear flow. We have made several

attempts to determine the direction of horizontal phase propaga-

tion. One attempt involved calculating the coherence between temp-

erature and horizontal velocity components as obtained by the pro-

filer in individual drops. In theory, observed phase differences

between temperature and east velocity component and temperature and

north velocity component, in the case of a single dominant wave,

would allow one to determine the direction of horizontal phase

propagation. In practice, the observed levels of coherence between

temperature and the two velocity components were too low to infer

anything about horizontal phase propagation direction. Using the

observed energy levels of the near-inertial waves in the main ther-

mocline and the mean vertical temperature gradient observed there,

a rough calculation indicates that a near-inertial wave with a

Page 136: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

136

frequency 20% greater than the local inertial frequency would pro-

duce a temperature signal that would be measurable in the profiles,

given the resolution level attainable by the EMVP temperature meas-

urements. The fact that the coherence between temperature and

velocity measurements is so low may be due to the presence of

temperature variations caused by waves with frequencies much greater

than the inertial frequency (the semidiurnal tide, for example).

In order to use the temperature profiles to determine horizontal

propagation direction, better frequency resolution of the profile

data must be obtained. Comparison of drops 235U and 236U (Figure

18), both visually and by doing lagged correlations between the

east components and north components of the two profiles, suggests

that strong features which can be identified between the two pro-

files occur at a slightly greater depth in profile 236U (4.8 kilo-

meters to the north of 235U). If this is true it implies that

these waves have a component of propagation toward the north.

Taking into account the fact that the mean shear flow is primarily

in the north-south direction (Figure 7), it would be expected

(Frankignoul ,1972) that a wave group with a component of propaga-

tion to the north and propagating energy downward would be acted

upon by the shear in such a way that the vertical wave length would

decrease with increasing depth. In fact, if the wave is thought to

be propagating directly north (against the major part of the mean

shear), an analysis of the WKB expressions for the change in verti-

cal wave length brought about by changing N and the presence of the

Page 137: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

137

shear suggests that the two influences could be of comparable mag-

nitude. However, Figure 3 shows that the most energetic waves have

wave lengths in the deep water that are longer relative to observed

wave lengths in the thermocline. We have also tried to use the

stretched, normalized profiles to see if there is a shift of ener-

getic spectral peaks in the upper versus lower parts of the

(stretched) water column. We chose a set of stretched profiles

(230D through 233D) which contained a distinct peak at about

130 sdb. Vertical wave number spectra for the upper and lower

halves of these profiles were then calculated separately. On the

assumption that the stretching of these profiles has removed

changes in vertical wave number and wave amplitude due to variations

of N, any change in the position of the peak in the upper half rela-

tive to the lower half might reflect the influence of shear. The

result showed that there is a slight shift toward higher stretched

vertical wave number in the lower half relative to the upper half,

but it is very small ( < 5 sdb). One might expect a shift in this

direction if the waves have a component of horizontal propagation

to the north. The fact that the shift is so small might indicate

that the waves are propagating primarily east-west, with a small

propagation component to the north.

The major conclusion of the above paragraphs is that a good

evaluation of the influence of the mean shear on wave propagation

awaits better estimates of horizontal phase propagation directions.

Page 138: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

138

A final important result of the present work is that wave

structures have been observed over the rough topography which have

polarization appropriate to upward energy propagation. As pointed

out in Chapter VI, there is a theoretical basis for believing that

rough bottom topography at the latitude of this experiment ( ~ 28°N)

could generate near-inertial waves that must propagate their energy

upward, away from the bottom. Bell (1975) has calculated that

rough topography could support an upward energy flux of - 1 erg/cm 2/

sec. If we accept our estimate of .6 - 1.0 erg/cm 2/sec downward

for the energy flux to internal waves input by some (as yet only

poorly understood) atmospheric forcing, then we see'that rough

topography generation of upward energy flux can be comparable to

atmospheric generation of downward energy flux. Some data from

current meters at 4000 meters in the MODE region (D. Porter, 1975,

personal communication) indicates that, in the band of periods from

37 hours to 14 hours (encompassing the inertial period), the total

energy in the frequency spectra obtained at 4000 meters depth over

the smooth topography is significantly less (at least by a factor

of 2) than the total energy in the above frequency band in spectra

obtained at the same depth over the rough topography. The above

observations suggest that, at least in regions of rough topography,

one might sometimes expect to see a convergence of energy toward

the thermocline, due to the combined effects of atmospheric and bot-

tom forcing. According to Frankignoul and Strait (1972), such a

convergence in the near-inertial frequency band can be seen in

Page 139: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

139

records obtained at Site D (390 20' N, 700 00' W). We again empha-

size that we have not seen any evidence of dominant upward energy

propagation (i.e. counterclockwise polarization with depth) in the

smooth topography time series.

The above discussion makes clear some of the problems asso-

ciated with interpreting the present analysis. These problems, in

turn, suggest some future experiments which could be performed in

order to clarify the influence of various processes on internal

wave propagation. We would now like to propose several relevant

experiments. First, a time series, similar to the one obtained

over smooth topography, should be performed over a rough topography

region. One of the primary difficulties with interpreting vertical

wave number spectra of deep water profiles over the rough topography

is that the low-frequency shear profile cannot be accurately removed

if only two profiles have been obtained at a given location (at a

time interval of one half of an inertial period). Any remaining

low-frequency components to the profiles will cause errors in the

calculated magnitudes of the clockwise and counterclockwise spectra.

Although the smoothed hodographs of Figures 30 and 31 are clearly

rotating counterclockwise with depth, and the main features in the

hodographs appear to reverse direction over the one-half inertial

period time interval between the two profiles, it is difficult to

put any confidence in the interpretation of spectra calculated from

single or paired profiles. A time series over rough topography

could result in other interesting observations as well. For

Page 140: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

140

example, one might speculate that the observed semidiurnal tidal

contribution to the profiles, which resides in roughly the lowest

five stretched vertical wave number estimates at the smooth topog-

raphy observation side, might show a change of character over the

rough topography. Interaction between the rough topography and

"low-mode" semidiurnal tidal waves might lead to the scattering of

the "low-mode" energy into higher modes. It would be interesting

in this regard to compare the DLRC calculations (Figures 24 to 27)

done for the smooth topography time series with the same calcula-

tion done for a time series over the rough topography. If, as

seems likely, the bottom-induced generation of near-inertial waves

is intermittent in time or space, then the "quick-look" capability

associated with the EMVP data processing system would be an in-

valuable aid in determining when and where to perform the rough

topography time series.

A second experiment involves using a pair of EMVP's to per-

form repeated paired profiles, such as those shown in Figure 18.

The single pair in Figure 18, as already pointed out, has shown

that there is an apparent tilt to energetic features that seem to

correspond to each other in the two profiles. Repeated paired

profiles would allow one to make a much more certain estimate of

this tilt and would also (if series of paired profiles were made

along two orthogonal directions) allow one to make an estimate of

the horizontal wave propagation direction.

_1~1_~ __j~_l~_~__~_~_

Page 141: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

141

A final possible experiment would involve the combined data

of an EMVP time series performed at some location (at the central

mooring, in smooth topography, for example) along with a current

meter mooring at the same location which is densely instrumented

for the first several hundred meters off the bottom. This exper-

iment would both help to verify the indication in the present EMVP

data that there is a layer near the bottom in which the average,

high-frequency, horizontal kinetic energy appears to decrease

toward the bottom and would also provide frequency information

which, as indicated in Chapter IV, is crucial in determining the

rate of energy loss undergone by near-inertial waves as they

reflect off the (smooth) bottom.

Perhaps the usefulness of a set of experimental observations

can best be measured by whether analysis of that set of data clearly

illuminates the way along which more comprehensive theoretical and

experimental work may be carried out. The present analysis has

generated both significant results and significant questions which

must be answered by further work, and, therefore, under the above

criterion this set of observational data has been extremely useful.

Page 142: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

a 142

APPENDIX A

Profile Locations and Times

Drop

177

178

179

180

181

182

183

184 t

185 t

186 t

187 t

188 t

189 t

190 t

191 t

192 t

193 t

194 t

195 t

196 t

197 t

Date 1973

April 10

" 10

" 10

" 12

" 12

" 12

May 14

" 15

" 15

" 15

" 16

" 16

" 16

" 17

" 17

" 17

" 17

" 18

" 18

" 18

" 18

Drop Time (GMT)

1627

1738

1858

1539

1658

1833

1458

1423

2045

2238

0430

1115

1750

0029

0547

1305

1842

0132

0801

1418

2044

Position

32021'.0N 64034'.5W

32021'.ON 64034'.5W

32021'.0N 64034'.5W

32021'.0N 64034'.5W

32021'.0N 64034'.5W

32021'.0N 64034'.5W

29024'.ON 69039'.8W

28001'.5N 68030'.OW

28001'.9N 68032'.8W

28002'.7N 68033'.8W

28001'.4N 68030'.3W

28003'.2N 68034'.0W

28002'.9N 68030'.9W

27056'.1N 68044'.5W

27057'.9N 68039'.8W

27056'.1N 68044'.3W

27058'.ON 68039'.OW

28008'.2N 68028'.3W

28"08'.0N 68018'.OW

28008'.5N 68028'.5W

28008'.6N 68018'.3W

Page 143: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

143

Drop

198 t

199

200

201

202

203

204

205

206

207

208 t

209 t

210

211

212

213

214

215

216

217

218

219 *

220 *

221 *

Date 1973

May 19

" 22

" 22

" 23

" 24

" 24

" 24

" 25

" 25

" 25

June 1

" 1

(" 8

1" 8

" 9

S" 9

5" 9

" 9

5" 9

" 10

" 10

" 11

" 11

" 11

Drop Time (Gt4T)

0231

0045

1612

1512

0737

1454

2034

0310

0915

2126

0204

0559

2257

2257

0530

1217

1217

1829

1829

1220

1220

0523

1152

1815

Position

28*02'.5N 68033'.6W

27006'.6N 69040'.12W

28000'.ON 69019'.OW

27043 ' . 2N 70001'.8W

27059'.3N 69039'.OW

27057'.ON 69036'.2W

27059'.2N 69038'.OW

27059'.1N 69038'.4W

27058'.ON 69038'.0W

29021'.7N 70042'.2W

23002'.ON 680 30'.8W

28002'.ON 68030'.9W

27023'.5N 70001'.2W

27023'.5N 70001'.2W

27029'.8N 7000'.4W

27023'.5N 70001'.2W

27023'.5N 70001'.2W

27029'.8N 70000'.5W

27029 '.8N 70000'.5W

27039 '.5N 69031'.0W

27039'.5N 69031'.0W

27059'.8N 69039'.1W

27059'.7N 69038'.9W

27059'.8N 69039'.2W

Page 144: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

Drop

222

223 *

224 *

225

226 *

227 *

228 *

229

230 *

231 *

232 *

233 *

234 *

235

236 *

237 *

238

239 *

240 *

241

242 *

243 *

244

245 *

Date 1973

June 11

" 12

" 12

" 12

" 12

" 13

" 13

" 13

" 13

" 13

" 14

" 14

" 14

" 14

" 14

" 14

" 14

" 14

" 15

" 15

" 15

" 15

" 15

" 15

Drop Time (GMT)

1815

0043

0714

1301

1945

0219

0759

1220

1800

2117

0102

0451

0828

1210

1206

1630

2019

2016

0034

0438

0438

0800

1219

1219

144

Position

27*59'.9N 69039'.0W

27059'.7N 69039'.3W

27059'.8N 69*39'.2W

27052'.0N 69039'.2W

27059'.8N 69039'.0W

27059'.8N 69039'.1W

27059'.8N 69039'.2W

27051'.6N 69042'.9W

27059'.7N 69039'.0W

27059'.7N 69038'.6W

27059'.8N 69038'.8W

27059'.8N 69039'.1W

27059'.8N 69038'.9W

27057'.6N 69038'.7W

27 059'.8N 69039'.IW

27059'.8N 69039'.1W

28003'.6N 69042'.0W

27059'.9N 69038'.8W

27059'.4N 69039'.1W

28005'.7N 69044'.7W

27059'.8N 69039'.1W

27059'.7N 69039'.1W

28000'.6N 69039'.8W

27059'.9N 69039'.0W

_____I*I~__/UYI*C___L~I^II~_.~~~II1~^- PL~I^-ZXI-

Page 145: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

Date 1973

June 15

" 15

Drop Time (GMT)

1935

2057

Position

27052'.1N 69039'.OW

27052'.2N 69038'.8W

(t) rough topography drops

(*) time series drops

246

247

145

_L________1__111_____Y___X ~_a~____ _YX~ I ~I___I___ ILI3*I__LU__.I

Page 146: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

146

APPENDIX B

Use of Electro-magnetic Velocity Profilers

to Measure Internal Waves

As was pointed out in Chapter I, a conversion equation must

be used to obtain horizontal water velocity from values of horizon-

tal electrical current density recorded by the EMVP. This conver-

sion equation involves not only the above variables, but also the

vertical water velocity at the point of measurement, as well as the

horizontal gradient of any electrical potential that exists in the

water. Vertical water velocities and horizontal potential gra-

dients, therefore, introduce errors into the calculation of hori-

zontal water velocity. We will now solve a rather simple problem

that allows us to obtain an estimate of this error.

The model chosen is that of a rotating ocean with constant

Brunt-Vaisalg frequency and constant depth, H, and with a rigid

surface and bottom. We assume that internal wave vertical modes

exist in the water column. This assumption is not crucial in esti-

mating the error, as will be seen later.

The equation for the electrical potential, 9, and the elec-

trical current density, S = (Jx, Jy, Jz), generated by an ocean

moving with velocity 9(x,y,z,t) = (G,vw) is:

A A

V$ x F -J/a , B-l

Page 147: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

147

where a is the conductivity of the sea water, which is assumed to

be constant, and t = (0,F y,Fz) is a vector representing the terres-

trial magnetic field. The coordinates x, y, and z are, respectively,

magnetic east, magnetic north, and vertical. The coordinate origin

is located at the sea surface, and z is positive upward.

Equation B-1 will be called the full conversion equation,

since it contains all terms that relate $, v and J. The horizontal

components of equation B-1 are:

Jx-=F Fz F -

a Z y ax

3-Y- -0iF - _a z ay B-2

The underlined terms make up the "applied" conversion equation, or

the equation that is actually used to convert EMVP electrical cur-

rent readings to horizontal water velocity. The remaining terms

cause the error. (The reader should refer to Sanford, et. al.

(1974) for a more comprehensive discussion of EMVP measurements.)

Since V * S = 0 and V x F = 0, the equation to be solved

becomes

V20 = V (v x F) = F (V x ) B-3

with the boundary conditions:

Page 148: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

148A

w = Jz = 0 at z = 0;

S= Jz = 0 at z = -H B-4

Evaluating x F, these boundary conditions may be rewritten:

W(z = 0) = w (z = -H) = 0 ;

~ (z = 0) = Fy * u(z = 0) ;az y

B-5AA (z = -H) = F ^u(z = -H) .

az y

Given (x,y,z,t) we can, at least in principle, solve for P and

then 3.

The internal wave equations are:

AU A

at _ f a2kat ax

axax

=- P+az

+ N2w = 0

v +aw+ -- +B-6

where u, v and w are, respectively, the east magnetic, north mag-

netic and vertical velocity components, f is the Coriolis parameter,

Page 149: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

149

and p is pressure divided by a reference density, o. We have used

the Boussinesq approximation, in which the density, ^ , is assumed

to be given by three terms:

P = Po + W(z) + p(x,y,z,t), B-7

and the water is assumed to be incompressible. The water pressure

is made up of two parts, pop and PoP. The hydrostatic part of the

pressure is given by:

P d(z) - g(P + (z)), B-8dz

where g is the acceleration due to gravity. Then poP is the devia-

tion of the pressure from hydrostatic, the bucyancy, b, is -gp/po

and the Brunt-Vais*l* frequency, N, is:

N= ( -B1 -9Sdz

By Fourier transforming equations B-6, where

i(kx + ly -wt)u(x,y,z,t) = u(z;k,l,w) e , B-10

and similarly for the other variables, we obtain the following

equation for w:

d2w + (k2 + 12) 2 2 = 0. B-idz2 W2 - f2

The form of w that satisfies the boundary conditions on vertical

velocity at the surface and bottom is:

Page 150: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

niTz

A sin Hn7T

H

With n an integer, and

n2r2 2 2 N2 )22 I (k2 + 12) N2 _ W

H2 w2 - f 2

Equations for u and v are:

N2 _ 2

iw(W2 _ f 2 )

N2 - 2

iW(W2 _ f 2 )

nrrzcos H

n7tH

cos H

nr -

H

(ilf + kw);B-14

(Iw - ikf)

B-15

- (k 2 + 12) F * (V x )

nirz

F sin HY nit

H

+F cos Hz nit

HH

(AN2 2 (ilf + kw) - ikA

iW(w2 _ f2)

Bf(k 2 + 12)

B = - -AL -2 2

nL iWm( 2 - f2 )Ti

150

B-12

B-13

Au -nit

H

AnTV

Then

dz2dz-

where

B-16

B-17

Page 151: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

151

The boundary conditions on are then:

S F ,B(ilf + kw)dz = n at z = 0 8-18dz nw

Hand

F yB(-l)n(ilf + kw)dz n at z = -H . B-19

H

The system of equations B-16, E-18 and B-19 can now be

solved to give:

Y=A N -~( ( i f - i k ) - i k+ n 2 + k2 (2 - f

SFy A(N2 -w 2 ) (if - ikw)

/f +72 W(W2 f2)( )t

x coth 42 + 12 H - (- 1 )n csch + 12 H x cosh 2 z

(A2 2+ 2 2( 2 f2 ) (If - ikw)-ikn k2 2 W(W2 f2)

FyA(N 2 _ 2) (If- ik) sinh X 2 +2 z

W(W2 _ f2)()2 V + 12

H

n7ir

FyA 2 ,2) (If - ikw) - ik

n2 2 +k 2 2 nH

- -H

Page 152: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

152nwz

F Af(N2 - 2) (k2 + 12) n'rzz Cos H+ i

- W(2 _ f 2 )HB-20

The full conversion equations have been given above (equa-

tions B-2). The applied conversion equations (those actually used

to convert from ) to F) are:

X (v -V) Fz , B-21

Jy -u )Fz

aB-22

where If and v are depth-independent terms (Sanford, 1971).

Since we are only interested in the part of J that varies with

depth, we set

= v =0.

If we now let utrue and vtrue be the velocity components astrue true

given by equations B-14 and B-15 and u and v beobs obs the compo-

nents calculated from the applied conversion equations B-21 and

B-22, then:

Utrue - Uobs = z 1 ay-Fz a y B-23

true obstrue obs 1F( + wF y)

We see that the expressions for 4 and w can be used to estimate the

error in the conversion equations. At a given depth, z, the energy

B-24

Page 153: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

153

in the u error signal is given by:

Eu (z) 1/2 < (U u- u u )* >error true obs true obs

2Fz ay B-25

where the brackets < > indicate an average over time. Similarly,

the v error signal is given by:

Ev (z) = 1/2 < ( v ) ( -v ) >

error true- obs true obs

2 < + w F 122Fz ax y B-26

We note that:

Eu (z) = 1/2 < u * u >true true true B-27

and

Ev (z) = 1/2 < * * v >true true true . B-28

Returning momentarily to the expression for 5 (equation

B-20) we see that the first two terms are proportional to

cosh /k2 + 12 z and (sinh (42 + 1 z)/ k2 + 12). We now make the

assumption that the horizontal component of the wave number vector,

S+ 12, is small enough so that.the cosh term is independent of z

and the sinh term becomes a linear function of z. Since we are only

Page 154: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

154

interested in that part of 4 that is depth-dependent, we ignore

the hyperbolic cosine term. The term (sini , T + 12 z)/( 2 + 12)

z as 2 ' 2i - O0. If we examine the coefficient multiplying

(sinh 2 + 12 z)/( 2 + 12) we see that for k2 + 12 << (n2 T2)/(H2)

the last term in the coefficient cancels the first term, and we are

left with (-i kF A)/( n ) as the coefficient of the term propor-y H2

tional to z. For k = 0 this term disappears. Even if k 0 this

term decreases rapidly as n increases; that is, for higher modes or

for frequencies close to f. Since we are only interested in those

terms that are sinusoidal in z, we will ignore the terms in ¢ that

are either independent of z or are only weakly dependent on z.

Then nrz

2:- 2 2 + k2 A 2FA (If - ikw) - ikj sin Hn + k 2 + 12 2) nH 2 +

-Y-- ~H

F Af(N 2 - W2) (k2 + 12) z---

n7 W(W2 _ f2) n_

H H B-29

(The reason for this assumption has been explained in Chapter II.

Any depth-independent contributions to the profiles examined in

Chapter II have been removed prior to analysis. Thus, the (approx-

imately) depth-independent contributions of 4 should not appear in

the profiles.)

We are particularly interested in the relative error in

uobs and v obs' given by Eu rror/Etrue and Ev /Everror true error true

Page 155: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

155

respectively. For simplicity, we have averaged the quantities

Euerror , Ev error, Eu , and Everror error true tru

Then

true H ) true0

over depth, z, from 0 to -H.

true > dz ,true

B-30

and similarly for the other three quantities. The final result is

that Eu /Eu and Ev /Everror true error true

Eue= rror

Tutrue

( 22 k

2

X k2 2 + 12f 2

( )1 HIf

F2n2 r2+z H2Hk2 + 12)

W2K2(W2 _ f 2 )

(N2 - 2) (k2 W2 + 12f2) +

are given by:

a2 - f2

F N2 - 2

F 2f2(k2 + 12)z

[t'' *lfZ(~ 2 I:(k2 + 2f2)( 27)2

H

Everror

Tv -Evtrue

x (12f 2 + k2W2

(H2n) 2

H

(n 2 +k2 + 12)2(12 + k2f2)

H

+ 2wk'( 2 - f2 ) + k2(W 2 - f)

(N2 _ 2) (N2 - W2)2

2W2 (W2 _ f 2 )2( + k2 + 12) F 2

(N2 -W2)2 F 2

( N 2 . 2 ) 2 F 2

(nH2 + k2 + 12)2w2(w2 - f 2 )H2

(N2 - W2)2

f2k2(k2 + 12)2+nT 2

It

B-31

F 2

B-32

h

Page 156: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

156

Considering first the u-component error, it can be shown

that Tu is a maximum, for a given w and n, if the wave is propa-

gating in the north-south direction (k = 0). Further, for waves

traveling in the east-west direction 1 = 0, and Tu = 0. For

1 n <<(n), we find that max 12

12 <(<()2, we find that max {Tu} 1 , if the frequency, w,nir 2

is close enough to f so that (w2 - f 2 )/(N 2 _ ,2) is small. Since

12 -2 _ f2

H 2 B-33

from equation B-13,

W2 _ f2max {Tu} a

N - 2 B-34

Equation B-33 or B-34 shows that the relative error in the u-compo-

nent energy depends only on frequency, w, or (what amounts to the

same thing) on the north-south "aspect ratio" of the wave, 1/(n/H).

Similar statements may be made about relative error in the

v-component energy. In this case, Tv is a maximum, for a given w

and n, if the wave is propagating in the east-west direction (1 = 0).

However, Tv does not go to zero as k goes to zero, because of the

presence of the term involving vertical velocity in the full conver-

sion equation for v. If (W2 _ f 2 )/(N2 _ 2) is small we find that

max (Tv} o((nw/H) B-35

Page 157: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

157

As was the case for Tu, max {Tv} for small (w2 - f2)/(N2 _ W2)

depends only on the frequency, or alternatively, on the east-west

aspect ratio k/(nr/H).

Using values of the Brunt-VMis&lN and inertial frequencies

appropriate to the MFODE center at a depth of 4500 dbar, we find

that the relative error terms max {Tu} and max {Tv} are both less

than 5% for periods down to about one-half day. This range of

periods from inertial down to one-half day contains roughly three

fourths of the energy observed in internal wave spectra from the

MODE area.

We have ignored vertical variations of N and the fact that

the observed waves are propagating in the vertical. Since the

error terms are proportional to (w2 - f2)/(N2 - w2 ), and this

would not change if waves were propagating vertically, the presence

of vertically propagating waves would not change the error estimate

significantly.

The influence of varying Brunt-Vaisal frequency is somewhat

more difficult to determine. If N varies weakly with z, this means

the forcing terms on the right hand side of equation B-17 have

coefficients that are weak functions of z. If we assume that N

varies on a scale that is appreciably longer than the scale of the

waves and try to solve the forced problem using two depth scales,

we would find that the lowest order estimate of the relative error

terms would again be proportional to (W2 - f 2 )/(N 2 _ W 2 ) but that

now N is a weak function of z. The value of N used above to get

Page 158: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

158

relative error estimates was taken to be that at 4500 dbar because

this value is close to the minimum average value in the N profile

and thus presumably maximizes the relative error.

Page 159: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

159

APPENDIX C

Accuracy and Precision of Observations

It is important to distinguish between accuracy and preci-

sion when interpreting velocity measurements obtained by the EMVP.

This is true because of certain characteristics of the EM1VP it-

self. A detailed description of the arrangement of sensors on the

EMVP can be found elsewhere (Sanford, et. al., 1974). It has been

pointed out (Sanford, 1971) that the electrodes that sense hori-

zontal electric current density only measure the depth-varying part

of the current. That is, there is a depth-independent part of the

current that is not sensed by the EMVP. Therefore, the instrument

measures horizontal current in much the same way that geostrophic

shear calculations do, when a "level of no motion" must be assumed.

This is clearly a loss of accuracy in the instrument.

A second cause of loss of accuracy in the EMVP involves the

actual location of the electrodes on the instrument. As the pro-

filer falls, it rotates. This rotation causes an oscillatory cur-

rent to flow in a coil within the profiler, since the coil is

rotating in the terrestrial magnetic field. This oscillating cur-

rent is then used to determine the point in a cycle of rotation

when the coil is pointing toward magnetic north. The orientations

of the two independent sets of electrodes on the profiler are de-

fined relative to the direction of this "compass coil." In the

present configuration of the instrument, one set of electrodes is

Page 160: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

160

oriented parallel to the compass coil, and one set is oriented

perpendicular to it. However, these "perpendicular" and "parallel"

orientations relative to the compass coil can only be measured to

about a degree. The result is that errors in defining the orien-

tation of the electrodes on the instrument cause errors in the

determination of the magnetic bearing of the horizontal electric

current density vector, and thus errors in the measured water ve-

locity. This is a systematic error, which is another cause of loss

of accuracy in the measurements. If, in a small section of a pro-

file, velocity readings from the two independent electrodes are

compared, it is usually found that although the two velocity read-

ings track each other closely with changing depth, there can be a

constant offset between the two of up to 1 cm/sec. If this offset

is removed, the actual RMS difference between the two readings is

usually less than 0.5 cm/sec. This RMS error can be taken to be

an estimate of the precision of the velocity measurements.

A more detailed discussion of EMVP velocity measurement

errors can be found in Sanford et. al. (1974). The above sources

of error should be kept in mind when interpreting velocity profiles.

As mentioned in the introduction, the EMVP also carries

pressure, temperature and conductivity sensors. We will consider

the pressure sensor error first. Because data in the EMVP is re-

corded in digital format, there is a quantizing error in any re-

corded variable. In the case of pressure, this quantizing error

amounts to about 1.5 dbar. Bottom pressure values as obtained by

Page 161: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

161

the EMVP for each drop in the time series have been compared in

order to estimate the precision of pressure measurements. If we

assume that the bottom at the time series site is flat, so that the

EMVP hit bottom at the same depth in each drop, then a comparison

of bottom pressures as recorded by the EMVP indicates a root-mean-

square error of about + 1 dbar at about 5500 dbar. This is a meas-

ure of the precision, or repeatability, of the pressure measurements.

On the other hand, accuracy is limited by the fact that values of

pressure computed from PGR (Precision Graphic Recorder) readings

could be determined to only about 5 dbar.

Temperature and conductivity measurements were notably less

successful than pressure measurements. The quantizing errors for

temperature and conductivity are 0.0170 C and 0.02 mmho/cm. The

quantizing error in temperature is twice what it should have been,

since the least significant bit in the digitizer was usually stuck

in the "1" mode. In addition, due to a failure of the conductivity

head on the profiler (the quartz liner in the head cracked under

high pressure and fell out), the conductivity measurements of the

head were seriously influenced by temperature and pressure. This

made the computation of salinity difficult. Although a curve was

derived which corrects conductivity for the above influences of

temperature and pressure on the head, the resulting computed salin-

ity was only accurate to about 0.035 o/oo. This was not judged to

be sufficient for density computations, and so these measurements

have not been used in the present work.

L~YYI___~___III__CIP___CII_____IY- .

Page 162: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

162

Finally, a comment should be made on navigational accuracy

during the time series. Ship positions were determined by Loran C.

Although nominally performed at one location, a comparison of re-

corded ship positions for all drops during the time series indicates

that the ship positions have a standard deviation of about + 230

meters (east-west) and t 270 meters (north-south) around a mean

position of 27* 59.75' N, 690 39.04' W. It should be noted that

the above numbers are based on ship position at the beginning of

each drop. The EMVP usually spent 5 to 10 minutes on the surface

before it began its descent. -During this time the instrument was

drifting free of the ship and was presumably carried to the south

by the mean flow (see Figure 5). This additional drift of the

instrument to the south would result in a shift of the mean posi-

tion of the time series to the south (probably by about 100 to 200

meters) but presumably would not increase the standard deviation of

the drop positions appreciably.

Page 163: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

163

APPENDIX D

A More General Treatment of the Partition of Energy

Between Clockwise and Counterclockwise Spectra

We let m be the vertical wave number in the stretched verti-

cal coordinate, f be the inertial frequency, and w be the wave fre-

quency. In what follows, we will use the vertical coordinate s,

where s = -z . Thus s is a stretched vertical coordinate increas-

ing upward. The horizontal velocity components of a single internal

wave may be represented by:

2ri(ms - t) y 2ni(ms - t)U (s,t;w,m) = e ; U (s,t;w,m) = -e , D-1L T W

where UL and UT are the horizontal velocity components parallel

and perpendicular to the direction of horizontal phase propagation,

respectively. Then (Garrett and Munk, 1972), the east and north

(u and v) velocity components are given by:

u = i UL cos - U sinO

and

v = i U sin + UT cos , D-2

where 4 is the horizontal phase propagation direction. We adopt

the convention that m is always positive. A wave with upward phase

propagation then has a positive frequency, w, while a wave with

downward phase propagation has a negative frequency. All

Page 164: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

164

contributions to the profiles from upward propagating waves may

then be represented by:N

SC 2ni(ms - at)UV (s,t;,) = ) A(w,m;) e dm dw

f 0

andN C

f (2 i(ms - wt)U (st,;) = A(w,m;4) e dm d D-3

7 D-3f 0

The amplitude function, A, is considered to be non-zero only for

f < w < N and m > 0. Therefore, the double integrals can be con-

sidered to be integrals from -o to +co. We similarly represent the

contributions to the profiles from downward propagating waves as:

0O 00

U0 (s,t;) = f (,,m;) e 2ni(msdm dw

and00 O

U0 (st;) f= B(w,m;o) e dm dwT ' D-4

In this case, B is non-zero only for -f < w < -N and m > 0.

The total horizontal velocity components are then given by:

uoU= U V+ U and U= U+ U . D-5

The cross correlation between u and v and the autocorrela-

tion for the u and v components are:

Page 165: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

165

p (.,t;)6l(T) -<< u(s,t)v*(s + ,t + T)

+ u (s,t)v(s + ?,t + T) 6(t)6(t - T) > D-6

and

[p (S,T;4) + p (-,T; )] 6(T)

= << u(s,t)u*(s + T,t + T) + u*(s,t)u(s + T,t + T)

+ v(s,t)v*(s + -,t + T) + v*(s,t)v(s + T,t + T) 6(t)6(t - T) > t>

D-7

where the angle brackets denote averages over s and t. The 6-func-

tions arise because we are making a linear section of the random

wave field along the s axis (Konyaev, 1973). (The contributions

to u and v at a given frequency and vertical wave number are

assumed to be a sum of waves with random phases.)

We assume that the above integral limits may be taken as - c

to + c. This implies an infinitely deep water column. Strictly

speaking, the limitation to a finite depth will cause the vertical

wave number spectra calculated below to be smoothed by a window

which is a function of (stretched) vertical wave number. The finite

depth limitation has been ignored in what follows. Using the above

expressions for the correlation functions, we can calculate the

cospectrum (P), quadrature spectrum (Q), and the total energy

spectrum (TE) for zero time lag. These quantities are given by:

Page 166: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

166

P(m;4) + iQ(m;4) = P (s,T; 4 )6 (T)e 2i(m ds dT

D-8and

TE(m;¢) =

-0 -00

(P, + P-27i(ms - wT) d

)(S-,T;4)6(T)e ds dT

D-9

Carrying out the required calculations, we find:

P(m;) + iQ(m;4) =

-N

+ f - f -

-f

f

1/4 IcossinOf

N

(1 - -)IA(w,m;c)j2dwW

jB(w,m; )12 dw

f IA(w,m;) 12 dw +

D-10

and

TE(m;) = 1/4 IA(m,w; )12 (1 +

f2(1 +(I + -- ) dw

D-1 1

We can simplify these expressions if we assume:

f2 ) dw

X~_ ~_~ __I(Il*_S11I____Y__Xitll^l ll~ - _1

f IB(o,m;)j'12 dW

JB(W,m;4) 12

Page 167: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

167

IA(m,w;)l 2 (1 - a(m,w'))E(w')A(m)R(4) ;

jB(m,w;4)j2 = a(m,w')E(w')A(m)R( ), D-12

where w' > 0. The value of a(m,w') is between zero and one. Use

of this coefficient allows us to indicate how much energy is con-

tributed to a one-sided frequency spectrum by waves with upward

and downward phase propagation. R(4) is assumed to be normalizedA

so that its integral around 2i is one. E(w') and A(m) could be

identified with the frequency and vertical wave number dependence,

respectively, of the Garrett-Munk spectrum given in Chapter V.

Integrating around 4, we get

N

TE(m) 1/4 E(')A(m) ( ,12 ) dw' 0-13

f

andN

Q(m) = 1/4 E(w')A(m) ( (1 - 2a) dw' D-14

f

The clockwise and counterclockwise spectra are given (to

within a constant) by:

C(m) = TE(m) + 2Q(m) ;

A(m) = TE(m) - 2Q(m) . (Gondlla, 1972). D-15

We point out that neither TE nor Q contains multiplicative factors

involving cos or sin4 in equations D-10 and D-11. These quantities

(and therefore C(m) and A(m) do not depend on the distribution of

~YY II-_ll-sX~~-l~lll__lI_ _ i-.-~I l^-LIIIIIC-~L~

Page 168: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

168

R(O) around a circle. This is .another way of saying that C(m)

and A(m) are invariant with respect to a rotation of the hori-

zontal coordinates. If we now assume that E(W') has a sharp peak

at a frequency w' near the inertial frequency, so that w'/f ~ 1,O 0

then

A(m) 1 - a(m, wo)

C(m) a(m, "4) D-16

which is the ratio of downward to upward energy propagation. For

a small a (which means that most of the energy is propagating down-

ward), A(m) is much larger than C(m). It is important to remember

that s is positive up. If we were to use z , the stretched verti-

cal pressure coordinate which increases downward (as for the spec-

trum in Figure 13, for example), then C(m), the clockwise spectrum,

would dominate. (Changing the vertical coordinate from s to z is

equivalent to inter-changing A and C.) Several other points can

be made. First, if a = 1/2 everywhere, the quadrature spectrum will

be zero, and clockwise and counterclockwise energy will be equal.

The choice of a = 1/2 corresponds to the case when, on the average,

equal energy is contained by waves with upward and downward energy

propagation. Second, if an energetic wave is present with a fre-

quency much greater than inertial (such as the semidiurnal tide),

we might expect that the addition of this wave will cause the ratio

A(m)/C(m) to approach one. It is interesting to note in this con-

text that the estimates of A(m)/C(m) (using z as the vertical

--̂ -L~plCUrYI11.~31~(P--LI I~-~--~Pq LI-~ -~- .I_.__I~LI~--~l

Page 169: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

169

coordinate) at the smallest vertical wave numbers in Figure 22

(where we know that the semidiurnal tide is important) do seem to

increase, relative to estimates at neighboring, higher stretched

vertical wave numbers.

----"i .~~~cll~ssr^l-------~~-rr~-u~-l *ayyu -

Page 170: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

170

BIBLIOGRAPHY

Bell, T. H., Jr., Topographically generated internal waves in theopen ocean, Jour. Geophys. Res., Vol. 80, No. 3, pp. 320-327, 1975.

Broglio, L., Review of Italian meteorological activities and re-sults, Proceedings of the First International Symposiumon Rocket and Satellite Meteorology, ed. by H. Wexler andJ. E. Caskey, Jr., North-Holland Publishing Co., Amsterdam,p. 94, 1963.

Charnock, H., A preliminary study of the directional spectrum ofshort period internal waves, Proc. 2nd U.S. Navy Symp. Mil.Oceanog., pp. 175-178, 1965.

Elford, W. G. and D. S. Robertson, Measurements of winds in theupper atmosphere by means of drfting meteor trails II,Jour. Atmos. Terr. Phys., Vol. 4, No. 5, pp. 271-284, 1953.

Endlich, R. M., R. C. Singleton and J.W. Kaufman, Spectral analysisof detailed vertical wind speed profiles, Jour. of the Atmos.Sciences, Vol. 26, No. 5, pp. 1030-1031, 1969.

Fofonoff, N. P., Spectral characteristics of internal waves inthe ocean, Deep-Sea Res., pp. 58-71, 1969.

Frankignoul, C. J., Stability of finite amplitude internal wavesin a shear flow, Geophys. Fluid Dynam., Vol. 4, No. 2,pp. 91-99, 1972.

Frankignoul, C. J. and E. J. Strait, Correspondence and verticalpropagation of the inertial-internal wave energy in thedeep sea, Mem. Soc. Roy. des Sci. de Liege, series 6,Vol. 4, pp. 151-161, 1972.

Garrett, C. and W. Munk, Space-time scales of internal waves,Geophys. Fluid Dynam., Vol. 2, pp. 225-264, 1972.

Garrett, C. and W. Munk, Space-time scales of internal waves: aprogress report, Jour. Geophys. Res., Vol. 80, No. 3,pp. 291-298, 1975.

I1~4PY ^Y"I" ~I*L NEWWWWWWWWON-~-

Page 171: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

171

Gonella, J., A totary-component method for analyzing meteorologicaland oceanographic vector time series, Deep-Sea Res., Vol. 19,No. 12, pp. 833-846, 1972.

Hauritz, B., H. Stommel.and W. H. Munk, On thermal unrest in theocean, Rossby Memorial Volume, New York: RockefellerInst. Press, pp. 74-94, 1959.

Hayes, S. P., Preliminary Measurements of the time-lagged coherenceof vertical temperature profiles, Jour. Geophys. Res.,Vol. 80, No. 3, pp. 307-311, 1975.

Hayes, S. P., T. M. Joyce and R. C. Millard, Measurements of verti-cal fine structure in the Sargasso Sea, Jour. Geophys.Res., Vol. 80, No. 3, pp. 314-319, 1975.

Hendershott, M. C.,"Inertial Oscillations of Tidal Period." Unpub-lished Ph.D. dissertation, Harvard University, 1964.

Hendry, R., Semidiurnal tides as observed in the MODE-1 field ex-periment, Draft Report of the MODE-I Dynamics Group, WoodsHole Oceanographic Institution. Unpublished manuscript,1974.

Hines, C. 0., Internal atmospheric gravity waves at ionosphericheights, Can. Jour. Phys., Vol. 38, No. 1, pp. 1441-1481,1960.

Hines, C. 0., Diurnal tide in the upper atmosphere, Jour. Geophys.Res., Vol. 71, No. 5, pp. 1453-1459, 1966.

Jenkins, C. M. and D. G. Watts, Spectral Analysis and Its Applica-tion. San Fransisco: Holden-Day, pp. 79-82, 1968.

Justus, C. G. and A. Woodrum, Upper atmospheric planetary-wave andgravity-wave observations, Jour. of the Atmos. Sciences,Vol. 30, No. 7, pp. 1267-1275, 1973.

Katz, E. J., Profile of an isopycnal surface in the main thermo-cline of the Sargasso Sea, Jour. Phys. Ocean., Vol. 3,No. 4, pp. 448-457, 1973.

Katz, E. J., Tow Spectra from MODE, Jour. Geophys. Res., Vol. 80,No. 9, pp. 1163-1167, 1975.

Konyaev, K. V. Spectral Analysis of Random Processes and Fields.Moscow: "Nayka" Publishing House. Translated by K.D. Leaman.Microform Publication MPOO01 from American Geophysical Union,Washington, D. C., pp. 129-132, 1973.

------i -Y~ ------C-'-~erY i-i-rrr~ i-~ iirarrsYICrrrYYIII~1 L ll~l~--~

Page 172: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

172

LaFond, E. C. and K. G. LaFond, Thermal structure through theCalifornia front, Report N.U.C. TP 224, p. 133, 1971.

Lazier, J. R. N., Temporal changes in some fresh water tempera-ture structures, Jour. Phys. Ocean., Vol. 3, No. 2,pp. 226-229, 1973.

Leaman, K. D. and T. B. Sanford, Vertical energy propagation ofinertial waves: a vector spectral analysis of velocityprofiles, Jour. Geophys. Res., Vol. 80 (in press) 1975.

McNary, J. F., Determination of current velocity profile usingan inclinometer line, University of Hliami Report,ML 68371, 90 pages, 1968.

Mooers, C. N. K., A technique for the cross spectrum analysis ofpairs of complex-valued time series, with emphasis onproperties of polarized components and rotational invariants,Deep-Sea Res., Vol. 20, No. 12, pp. 1129-1141, 1973.

Muller, P. and D. J. Olbers, On the dynamics of internal waves inthe deep ocean, Jour. Geophys. Res., (in press) 1975.

Munk, W. and N. Phillips, Coherence and band structure of inertialmotion in the sea, Rev. Geophys., Vol. 6, No. 4, pp. 447-471, 1968.

Phillips, 0. M., Energy transfer in rotating fluids by reflectionof inertial waves, Phys. Fluids, Vol. 6, No. 4, pp. 513-520, 1963.

Phillips, 0. M. The Dynamics of the Upper Ocean. Cambridge, England:Cambridge University Press, pp. 173-174, 1966.

Pochapsky, T. E., Internal waves and turbulence in the deepocean, Jour. Phys. Ocean., Vol. 2, No. 1, pp. 96-103, 1972.

Pochapsky, T. E. and F. D. Malone, A vertical profile of deephorizontal current near Cape Lookout, North Carolina, Jour.Mar. Res., Vol. 30, No. 2, pp. 163-167, 1972.

Pollard, R. T., On the generation by winds of inertial waves inthe ocean, Deep-Sea Res., Vol. 17, No. 4, pp. 795-812, 1970.

Rossby, H. T., A vertical profile of currents near PlantagenetBank, Deep-Sea Res., Vol. 16, No. 4, pp. 377-385, 1969.

Page 173: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

173

Sanford, T. B., Motionally induced electric and magnetic fieldsin the sea, Jour. Geophys. Res., Vol. 76, No. 15, pp. 3476-3492, 1971.

Sanford, T. B., R. G. Drever and J. H. Dunlap, The design andperformance of a free-fall electro-magnetic velocityprofiler (EMVP). Woods Hole Oceanographic Institutionreference 74-76. Unpublished manuscript. 1974.

Smith, L. B., The measurement of winds between 100,000 and300,000 feet by use of chaff rockets, Jour. Meteor., Vol.17, No. 3, pp. 296-310, 1960.

Titov, V. B. and L. M. Fomin, Vertical structure of currents inthe Indian Ocean based on measurements, Oceanology , Vol.11, No. 4, pp. 486-493, 1971.

Voorhis, A. D., Measurements of vertical motion and the partitionof energy in the New England slope water, Deep-Sea Res.,Vol. 15, No. 5, pp. 599-608, 1968.

Voorhis, A. D. and R. R. Benoit, MODE SOFAR Float in situ datasummary, Woods Hole Oceanographic Institution TechnicalReport 74-37. Unpublished manuscript. 1974.

Webster, F., Observations of inertial-period motions in the deepsea, Rev. Geophys., Vol. 6, No. 4, pp. 473-490, 1968.

Webster, F., Turbulence spectra in the ocean, Deep-Sea Res.,Suppl. to Vol. 16, pp. 357-368, 1969.

Wunsch, C., Progressive internal waves on slopes, Jour. FluidMech., Vol. 35, No. 1, pp. 131-144, 1969.

Yih, C.-S. Fluid Mechanics. New-York: McGraw-Hill. p. 44, 1969.

_LL--I-I 11.I-I^^^Il X~E~IY

Page 174: THE VERTICAL PROPAGATION OF INERTIAL WAVES IN THE …

174

BIOGRAPHICAL NOTE

NAME: Kevin Douglas Leaman

BORN: November 23, 1948, Lakewood, Ohio

MARITAL STATUS: Married to the former Diane Saul

EDUCATION:

Institution

Univ. ofWashington

Univ. ofMichigan

EMPLOYMENT:

Institution

Univ. ofMichigan

Woods HoleOcean. Inst.

Dates ofAttendance

9/66-6/67

9/67-6/71

Dates ofEmployment

9/69-6/70

1/74-6/75

DegreeEarned

B.S.M.S.

Date ofDegree

12/7012/72

Major Field

Oceanography

OceanographyOceanography

Duties

Teaching Assistant in Departmentof Meteorology and Oceanography

Research Assistant inPhysical Oceanography

PUBLICATIONS:

Translation from Russian to English of Konyaev, K. V. SpectralAnalysis of Random Processes and Fields. Moscow: "Nayka" Pub-lishing House. Microform Publication MP0001 from AmericanGeophysical Union, Washington, D. C., 1973.

Leaman, K. D. and T. B. Sanford, Vertical energy propagation ofinertial waves: a vector spectral analysis of velocity profiles,Jour. Geophys. Res., Vol. 80 (in press), 1975.